paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1208.5513 | 1 | 1208 | 2012-08-27T21:38:15 | The growth and hydrodynamic collapse of a protoplanet envelope | [
"astro-ph.EP"
] | We have conducted three-dimensional self-gravitating radiation hydrodynamical models of gas accretion onto high mass cores (15-33 Earth masses) over hundreds of orbits. Of these models, one case accretes more than a third of a Jupiter mass of gas, before eventually undergoing a hydrodynamic collapse. This collapse causes the density near the core to increase by more than an order of magnitude, and the outer envelope to evolve into a circumplanetary disc. A small reduction in the mass within the Hill radius (R_H) accompanies this collapse as a shock propagates outwards. This collapse leads to a new hydrostatic equilibrium for the protoplanetary envelope, at which point 97 per cent of the mass contained within the Hill radius is within the inner 0.03 R_H which had previously contained less than 40 per cent. Following this collapse the protoplanet resumes accretion at its prior rate. The net flow of mass towards this dense protoplanet is predominantly from high latitudes, whilst at the outer edge of the circumplanetary disc there is net outflow of gas along the midplane. We also find a turnover of gas deep within the bound envelope that may be caused by the establishment of convection cells. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 23 January 2021
(MN LATEX style file v2.2)
The growth and hydrodynamic collapse of a protoplanet envelope
Ben A. Ayliffe1,2(cid:63) and Matthew R. Bate1
1School of Physics, University of Exeter, Stocker Road, Exeter EX4 4QL
2Monash Centre for Astrophysics (MoCA) & School of Mathematical Sciences, Monash University, Clayton, Vic 3800, Australia
23 January 2021
ABSTRACT
We have conducted three-dimensional self-gravitating radiation hydrodynamical models of
gas accretion onto high mass cores (15-33 M⊕) over hundreds of orbits. Of these models, one
case accretes more than a third of a Jupiter mass of gas, before eventually undergoing a hy-
drodynamic collapse. This collapse causes the density near the core to increase by more than
an order of magnitude, and the outer envelope to evolve into a circumplanetary disc. A small
reduction in the mass within the Hill radius (RH) accompanies this collapse as a shock prop-
agates outwards. This collapse leads to a new hydrostatic equilibrium for the protoplanetary
envelope, at which point 97 per cent of the mass contained within the Hill radius is within the
inner 0.03 RH which had previously contained less than 40 per cent. Following this collapse
the protoplanet resumes accretion at its prior rate. The net flow of mass towards this dense
protoplanet is predominantly from high latitudes, whilst at the outer edge of the circumplan-
etary disc there is net outflow of gas along the midplane. We also find a turnover of gas deep
within the bound envelope that may be caused by the establishment of convection cells.
Key words: planets and satellites: formation -- methods: numerical -- hydrodynamics -- radia-
tive transfer
1
INTRODUCTION
The ideas discussed in this paper begin with Perri & Cameron
(1974), who stated that "when the mass of the core becomes suf-
ficiently great, the surrounding gaseous envelope will become hy-
drodynamically unstable against collapse onto the planetary core".
This process is controlled by the battle between gravity that acts to
contract an envelope onto the core and the gas pressure which acts
to support it.
Mizuno (1980) performed stability calculations to determine
the combinations of core masses and opacities that would make
a protoplanetary envelope unstable to collapse. Work of his con-
temporaries considering the structure of giant planets in the solar
system suggested that each such planet (Jupiter, Saturn, Uranus,
& Neptune) possessed a solid core with a mass of order 10 M⊕
(Slattery 1977; Hubbard & Macfarlane 1980). Using these values
as a target, and assuming a fixed accretion rate, Mizuno concluded
that a grain opacity of κ ≈ 1 cm2 g−1 was required in the enve-
lope material during formation to trigger a collapse when the core
mass was ≈ 10 M⊕. Lower opacities were found to lead to enve-
lope collapse at lower core masses. Following the envelope col-
lapse Mizuno states that continued accretion is likely required for
the protoplanet's to attain their final masses.
It was suggested later by Bodenheimer & Pollack (1986), who
performed evolution calculations that included evolution beyond
the attainment of the critical core mass, that rather than a dynam-
(cid:63) E-mail:[email protected]
© 0000 RAS
ical collapse, the envelope may quasi-statically contract onto the
solid core. They suggested that if a sufficient mass of molecular
hydrogen in the envelope was apt to undergo dissociation, then this
would remove enough energy from the contraction to bring about a
dynamical collapse. However, their models indicate that such disso-
ciative regions possess an insufficient fraction of the envelope mass
for this to occur.
In the early 1990s Wuchterl wrote a series of papers exploring
the evolution of a protoplanetary envelope through the hydrostatic
phases approaching the critical core mass, and in what he found to
be a subsequent hydrodynamic phase (Wuchterl 1991). Solving the
equations of radiation hydrodynamics in one-dimension, Wuchterl
indicated that following a period of quasi-static contraction, during
which the envelope heats up, the transport of this heat out through
the envelope by convection and radiation perturbs the hydrostatic
equilibrium. In particular Wuchterl cites the κ − mechanism as the
means of exciting the dynamical waves that destabilise the enve-
lope. The result of this hydrodynamic evolution was the ejection of
a large fraction of the envelope mass, rather than an inwards col-
lapse as had previously been supposed by Perri & Cameron (1974).
Pollack et al. (1996) continued performing models assuming a
quasi-hydrostatic contraction, and suggested that further work was
required to consider the hydrodynamic evolution of young proto-
planets to establish the proper evolution scenario. A step in this di-
rection was taken by Tajima & Nakagawa (1997) who performed a
stability analysis of a growing protoplanetary envelope using a dis-
tinct numerical code to that of previous authors. Their aim was to
determine whether quasi-static contraction, or an envelope instabil-
2
B.A. Ayliffe & M.R. Bate
ity akin to that suggested by Wuchterl, was the more likely evolu-
tionary course for a growing protoplanet. They perturbed the enve-
lope at intervals during its evolution to see if such action might push
a marginally stable system towards instability. They concluded that
quasi-static contraction was viable for a protoplanet growing all the
way to a Jupiter mass.
There has since been a substantial amount of work consid-
ering the gas accretion rates that cores might achieve in circum-
stellar discs with a variety of properties. Ikoma et al. (2000) per-
formed quasi-static evolutionary models to determine the depen-
dence of gas accretion upon core mass, grain opacity, and the core's
accretion history, finding that these factors were strongly, mod-
erately, and weakly significant respectively. Bryden et al. (1999)
and Lubow et al. (1999) performed locally-isothermal hydrody-
namics simulations of discs containing planetary cores, the latter
finding that the accretion rate drops off as the protoplanet mass be-
comes very large; a result of the broadening disc gap that it forms.
Further hydrodynamic models with more realistic thermodynamics
were performed by D'Angelo et al. (2003a), Klahr & Kley (2006),
Paardekooper & Mellema (2008), and Ayliffe & Bate (2009a), find-
ing similar turn overs in the accretion rate with increasing mass, and
illustrating the impact of grain opacity on accretion. These models
have generally had limited resolution in the vicinity of the proto-
planet, and though Ayliffe & Bate (2009a) achieved high resolu-
tion, the evolutionary period was extremely short due to the com-
putational demands of the calculations.
In this paper we report results from three-dimensional self-
gravitating radiation hydrodynamics calculations that resolve the
protoplanet's envelope, whilst modelling its hydrodynamic evolu-
tion and growth within a section of a circumstellar disc. Using high
mass discs (though still stable; Toomre Q >> 1), and assuming low
opacities, we achieve accretion rates that allow significant envelope
growth in only a few hundred orbits, allowing us to examine their
development. Our computational method is described in Section 2,
followed by our results in Section 3, and a discussion of their rela-
tionship with previous results in Section 4.
2 COMPUTATIONAL METHOD
The calculations discussed in this paper were performed using a
three-dimensional SPH code. This SPH code is derived from a code
first developed by Benz (1990; Benz et al. 1990) which has un-
dergone substantial modification in subsequent years. Energy and
entropy are conserved to timestepping accuracy by use of the vari-
able smoothing length formalism of Springel & Hernquist (2002)
and Monaghan (2002), where our particular implementation is de-
scribed in Price & Monaghan (2007). Gravitational forces are cal-
culated and particle neighbours are found using a binary tree. Ra-
diative transfer is modelled using the flux-limited diffusion approx-
imation, employing the method developed by Whitehouse, Bate, &
Monaghan (2005) and Whitehouse & Bate (2006). Integration of
the SPH equations is achieved using a second-order Runge-Kutta-
Fehlberg integrator with particles having individual timesteps (Bate
1995). Gas within the models is subject to an artificial viscosity, im-
plemented in a parameterised form as developed for SPH by Mon-
aghan & Gingold (1983), and modified to deal with high Mach
number shocks by Monaghan (1992). The code has been paral-
lelised by M. Bate using OpenMP and MPI.
2.1 Model setup
The calculations we have performed were conducted by modelling
a small section of a circumstellar disc, centred upon a protoplane-
tary core and corotating with its orbit. The protoplanet is modelled
as a gravitating mass with a 'surface'. It attracts gas from the disc,
building up an atmosphere which is supported by the surface. The
disc section measures r = 1±0.15 rp (5.2±0.78 AU), and φ = ± 0.15
radians. The protoplanet is sited at a radius of rp, which in all cases
is equivalent to 5.2 AU, and orbits about a star of 1 M(cid:12). The disc
has a surface density profile of Σ ∝ r−1/2, and a temperature profile
of Tg ∝ r−1 (giving a scaleheight of H/r = 0.05), where these pro-
files are equivalent to those used in our previous work, and were
originally chosen to match work of Lubow et al. (1999) and Bate
et al. (2003). The initial temperature at rp is ≈ 73 K, where this tem-
perature is taken to principally result from stellar irradiation. Near
the midplane of the disc, this initial temperature may increase due
to viscous heating as the model evolves if the opacity is sufficient
to insulate the region against rapid radiative cooling. However, at
the upper and lower radiation boundaries (see section 2.3 below),
where the medium is always optically thin, an assumption of tem-
peratures dominated by the stellar irradiaton is reasonable. The sur-
face density at rp has an unperturbed value of 750 g cm−2, which
gives a relatively massive disc, comprising 0.1 M(cid:12) of gas within
25 AU. We use a high disc surface density because Ayliffe & Bate
(2009a) found that increasing the disc surface density resulted in
somewhat faster accretion rates onto embedded protoplanets and
the goal of this paper is to investigate the three-dimensional evolu-
tion of protoplanets that accrete massive gaseous envelopes.
Further details of our model setup are given below, but the
method is identical to that employed in Ayliffe & Bate (2009a),
and that paper contains somewhat more extensive information, in-
cluding resolution tests. We note that the most interesting model
discussed in this paper, a case which results in a hydrodynamic
envelope collapse, required more than 6 CPU years to reach its fi-
nal state. It is this extensive calculation time which has limited the
number of models that is has been possible to perform.
2.2 Disc sections
Within the disc section being modelled, the particles are initially
distributed according to the underlying density profile, and their
velocities are Keplerian. The protoplanet is orbiting its star in an
anticlockwise direction, and in the corotating frame of the mod-
elled section this leads particles at r < rp to orbit anticlockwise, and
those at r > rp to orbit in a clockwise fashion. Particles encounter-
ing the boundary of the section are removed from the calculation,
whilst a number of ghost particles beyond the domain of the calcu-
lation, along its boundaries, act to replicate the pressure and viscous
forces expected from a continuous disc. To prevent the depletion of
gas within the disc section, particles are injected along the bound-
aries where the material should flow in. For the anticlockwise orbit-
ing gas this input is along the φ = −0.15 radian boundary, between
r = 0.85− 1.0 rp, whilst the clockwise flowing gas is injected along
the φ = 0.15 radian boundary, between r = 1.0 − 1.15 rp. The in-
jection scheme does more than simply replace gas which leaves the
section. The velocity and density structure of the injected gas is ob-
tained from three-dimensional global simulations of protoplanets
embedded in discs performed using ZEUS (Bate et al. 2003), such
that a gap is opened in the disc which corresponds to the mass of
the embedded protoplanet. The models of Bate et al. used a sur-
face density of 75 g cm−2, but the particles injected in the calcula-
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
3
tions discussed here have their masses scaled to reflect the chosen
surface density of 750 g cm−2. As the protoplanet grows the gap
width is suitably increased by interpolation through the protoplanet
mass range provided by Bate et al. (2003) (1 M⊕−1 MJupiter). The
injected particles come to dominate the section after less than five
orbits, ensuring the structure is consistent with the presence of the
protoplanet.
Model
A
B
C
D
E
J
Core mass
[M(cid:12)]
4.54 × 10−5
6 × 10−5
6 × 10−5
6 × 10−5
6 × 10−5
1 × 10−4
[M⊕]
15
20
20
20
20
33
Core radius × Physical Opacity
[% IGO]
core size
[rp]
2.2 × 10−4
6.6 × 10−5
6.6 × 10−5
2.2 × 10−4
2.2 × 10−4
2.54 × 10−4
10
3
3
10
10
10
0.1
0.1
1
0.1
1
0.1
2.3 Thermodynamics
All the calculations discussed in this paper were performed using
radiation hydrodynamics, employing a two temperature (gas and
radiation) radiative transfer scheme using a flux-limited diffusion
approximation (as described by Whitehouse et al. 2005 and White-
house & Bate 2006). Work and artificial viscosity act to increase
the thermal energy of the gas, and work done on the radiation field
increases the radiative energy, which can be transported via flux-
limited diffusion. The energy transfer between the gas and radiation
fields is dependent upon their relative temperatures, the gas density,
and the gas opacity.
The gas is treated using an ideal gas equation of state p =
ρTgRg/µ where Rg is the gas constant, ρ is the density, Tg is the gas
temperature, and µ is the mean molecular mass. The thermal evolu-
tion takes into account the translational, rotational, and vibrational
degrees of freedom of molecular hydrogen (assuming a 3:1 mix of
ortho- and para-hydrogen; see Boley et al. 2007). Also included are
molecular hydrogen dissociation, and the ionisations of hydrogen
and helium. The hydrogen and helium mass fractions are X = 0.70
and Y = 0.28, respectively, whilst the contribution of metals to the
equation of state and the thermal evolution is neglected.
The flux-limited diffusion scheme transfers energy between
SPH particles, which does not enable it to radiate into a vacuum.
In order that the disc can cool from its upper and lower surfaces,
a boundary is applied that maintains the initial temperature profile
in the high atmosphere of the disc. This boundary is situated at a
height above/below the midplane that corresponds to the edge of
the optically thick region, that is where the optical depth (τ) from
outside the disc to that depth is τ ≈ 1. SPH particles comprising
the boundary regions evolve normally, but their energies are set
according to the initial radial profile, allowing them to act as energy
sinks.
2.4 Opacity treatment
We use the opacity tables of Pollack et al. (1985) to provide grain
opacities, whilst the tables of Alexander (1975) (the IVa King
model) provide the gas opacities at higher temperatures when the
grains have sublimated. The former table gives interstellar grain
opacities (IGO) for solar metallicity molecular gas, but in this work
we reduce these opacities by orders of magnitude below this nomi-
nal level; we divide by factors of 100 and 1000. The justification for
this comes of the likely agglomeration or sublimation of grains in
the vicinity of a forming protoplanet (Podolak 2003; Movshovitz
et al. 2010). We do not modify the gas opacities, but enforce a
minimum for the grain opacities at the interface between the two
regimes that corresponds to the gas minimum ensuring a smooth
transition (see Ayliffe & Bate 2009a for more details).
© 0000 RAS, MNRAS 000, 000 -- 000
Table 1. Properties of the various models described in this paper, all of
which were performed in a disc with a surface density of 750 g cm−2 at rp.
The opacity is given as a percentage of the interstellar grain opacity (IGO)
assumed in the opacity tables we employ. Core radii are all based on the
models of Seager et al. 2007, multiplied by factors of 3 and 10 as marked.
2.5 Planetary Cores
The planetary cores in these simulations are modelled by a grav-
itational potential, and a surface potential that yields an opposing
force upon gas within one core radius of the core's surface. The
combination of the gravitational and surface forces takes the form
of a modification to the usual gravitational force as
1 −
Fr = − GMc
r2
(cid:32) 2Rc − r
Rc
(cid:33)4
(1)
for r < 2 Rc where r is the radius from the centre of the plane-
tary core, Rc is the radius of the core, and Mc is the mass of the
core (see Ayliffe & Bate 2009a for further details). This equation
yields zero net force between a particle and the planetary core at
the surface radius Rc, whilst inside of the core's radius the force is
outwards and increases rapidly with decreasing radius. Gas parti-
cles therefore come to rest very close to the core radius, though the
equilibrium position is slightly inward of this value due to the pres-
sure exerted by the gas that accumulates on top of the inner most
layer of particles.
Seager et al. (2007) calculate core radii for solid exoplanets,
amongst which are cases comprising of 75 per cent water, 22 per
cent silicates, and 3 per cent iron. We use these models to deter-
mine the realistic sizes of protoplanetary cores that correspond to
the masses used in this paper. These core radii were employed for
a number of calculations in Ayliffe & Bate (2009a), but these cases
were not evolved for many orbits due to the short timesteps required
deep within the planetary potential. To follow the evolution over
longer periods it has been necessary to perform calculations using
larger core radii. To this end we scale up these previously used core
radii by factors of 3 or 10 to reduce the required computation time.
We also performed new models using the realistic core radii, but
these calculations were too slow to give useful results and thus are
not reported here. The properties of our different models are given
in Table 1. The rate at which the planetary core accretes gas may
be affected by the form of the surface potential described by equa-
tion 1, and this is explored briefly in Section 3.1. A smooth start to
the calculations is provided by shrinking exponentially towards the
desired Rc from an initial radius of 0.01 rp over the course of the
first orbit.
4
B.A. Ayliffe & M.R. Bate
Figure 1. Accretion histories for protoplanets with initial core masses of
15 (A), 20 (D, E), and 33 ( J) M⊕. The models possess core radii 10 times
their likely physical size. Model E is performed using a disc with 1 per cent
interstellar grain opacity, whilst all the others use 0.1 per cent; model E
is otherwise identical to D. The size and opacity dependencies are further
illustrated in Fig. 2.
2.6 Measuring the gas accretion rates onto the planetary
cores
We measure the gas accretion rates by calculating the rate at which
gas passes into the self-consistently calculated Hill sphere of the
protoplanet given by
(cid:115)
RH = 3
Mp
3M∗
rp
(2)
where Mp is the protoplanet mass which is the sum of the core mass
(Mc) and the accreted mass (Macc), where accreted mass comprises
all the gas within RH. The gas mass is discretised amongst the SPH
particles, allowing iteration through equation 2 until such time as
the addition of a particle's mass to Macc no longer increases RH
sufficiently to encompass the next available particle.
The net flux through the Hill radius corresponds to the growth
rate of the envelope and/or circumplanetary disc, which are the only
repositories for gas that fails to remerge from this region. Our use
of the Hill radius to measure the mass growth is arbitrary, but rea-
sonable, since any protoplanet must be smaller than the Hill radius.
It is important to note that the Hill radius does not define the extent
of the envelope, which tends to be smaller (e.g. ∼ 0.25 RH Lissauer
et al. 2009). However, the difference between the mass accretion
rates (and total accreted masses) as measured at the Hill radius and
at 0.3 RH is small except at the start of the calculation (Section 3.2).
3 RESULTS
We have performed three-dimensional radiation hydrodynamics
models of the accretion of gas onto planetary cores (or embryos)
with a range of masses, over hundreds of orbits, in discs of varying
Figure 2. Accretion history for a 20 M⊕ core with 0.1 per cent and 1 per cent
opacities (as marked), for two different core radii, 3× physical (dashed lines,
models B & C), and 10× physical (solid lines, models D & E). Note that for
the higher opacity case, which accretes more slowly, the models of different
assumed core radii coevolve. For the lower opacity cases, which accrete
large envelopes much more rapidly, the case with the smaller core radius
accretes more slowly that the large core case during the initial ∼ 50 orbits.
However, thereafter, when their accretion rates are measured at equivalent
masses, they exhibit very similar rates.
opacity. This work extends upon models we performed in Ayliffe &
Bate (2009a) where we followed the accretion for a relatively short
period. Moreover, by using discs with highly reduced opacities, the
models presented here include the accretion of much more signif-
icant envelope masses. In one case the accreted mass is sufficient
to trigger a hydrodynamic collapse of the envelope (Section 3.2),
followed by a return to steady gas accretion.
3.1 Envelope accretion
We performed calculations starting with various core masses rang-
ing from 15 − 33 M⊕ (Table 1). The principal property that controls
the accretion rate of a protoplanet of a given mass is the opacity of
its envelope. A lower opacity enables more rapid radiative cooling,
allowing the envelope to contract more quickly and so accrete gas
at a faster rate (Hubickyj et al. 2005; Papaloizou & Nelson 2005;
Ayliffe & Bate 2009a). The models presented in this work exploit
this dependence to accelerate the growth process by adopting re-
duced opacities. The impact of opacity on the growth rate of a pro-
toplanet is demonstrated by comparing Models E and D in Figs. 1
and 2. Model E employs an opacity ten times larger than D, which
results in a much slower rate of accretion for the former.
In each case we introduced a bare core with no preexisting
envelope. This results in a rapid initial gas accretion rate to form
a quasi-static envelope, followed by a period of slowing accre-
tion. This initial growth phase is not realistic for a protoplanet
that forms in situ and concurrently accretes both solids and gas
during the core formation phase (e.g. Pollack et al. 1996; Alibert
et al. 2005). In models that account for both the core and enve-
lope growth, the gas accretion rate tends to have an extended pe-
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
5
this is evident in Fig. 3 which illustrates that the fit is still extremely
good at 160 orbits. Fitting Model J over the same 40 orbits, where
an exponent b = 0.45 is found to be best, the trend of an initially
reducing rate of accretion falters at the limit of the fitting region,
around 50 orbits. At this point the accreted mass is a little more
than half the core mass, and gravity begins to dominate the growth,
leading to accelerating accretion.
In all cases, such as the models shown in Fig. 1, the transition
from a slowing rate of accretion to an accelerating rate of accretion
occurs once the envelope mass has reached half the core mass. This
begins the process of runaway accretion, that continues until such
time as the envelope potentially undergoes a dynamic collapse, as
will be discussed in the following section. It is this period of accre-
tion that leads to the most significant growth of the protoplanet's
mass.
As mentioned in Section 2.5, in order to make these three-
dimensional models computationally viable, we were forced to
adopt non-realistic core radii for the planetary surfaces. In Ayliffe
& Bate (2009a) we compared the accretion rates achieved in mod-
els using different core radii and found that for high opacities the
accretion rates did not depend significantly on the core radius that
was used, but for low opacities (1 per cent and 0.1 per cent interstel-
lar grain opacities) the accretion rates obtained with smaller cores
were significantly lower than for equivalent models adopting larger
cores. However, the earlier models were only evolved for 10 orbits,
which is still during the initial phase of envelope creation when the
accretion rate is decreasing. Fig. 2 illustrates the growth of 20 M⊕
cores embedded in a disc of either 0.1 or 1 per cent interstellar grain
opacity (as marked). The solid lines present models with planetary
cores 10 times the realistic core radii, and the dashed lines 3 times.
Over the course of the initial settling period the different core radii
in the 0.1 per cent opacity models lead to some divergence in the
growth, as is evident. However, measuring the accretion rates of
the two calculations at equivalent masses beyond the initial 50 or-
bits, it is found that these rates deviate by no more than 20 per cent
from one another. In the 1 per cent opacity case, the results ob-
tained using the two different core radii are indistinguishable from
each other. We believe that measuring the accretion rates at longer
times, when the models are more established is responsible for the
relatively small differences seen with different core radii. However
in this case we are only varying the core radius by just over a factor
of 3, whilst in Ayliffe & Bate (2009a) the comparison was made
spanning more than a factor of 12. At the present time we do not
have any models with which we can ascertain to what extent the
accretion rate differences measured in this previous work were due
to the large factor difference in the core radius, and what fraction
came of the early times at which the measurements were made.
As such, the enlarged radii of the cores that we have adopted here
should be kept in mind.
In the rest of this paper we focus on one particular case, Model
J, that accretes a very significant envelope in a short period of
time that undergoes a dynamic collapse. The other calculations dis-
cussed in this section are ongoing, and will eventually allow us to
explore the evolution of envelopes that are built up under less ex-
treme conditions (i.e. with slower accretion rates).
3.2 Hydrodynamical collapse - Model J
Model J accretes the most significant mass of any of our models, as
is to be expected given its favourable conditions. The 33 M⊕ core is
the most massive that we employ, and in this case is coupled with
a large 10 times realistic core radius, and a 0.1 per cent interstellar
Figure 3. A comparison between results from the lower surface density,
shorter evolution calculations presented in Ayliffe & Bate (2009a) and the
longer, more rapidly accreting models discussed in this paper. The lower
solid line is the accretion history over 160 orbits for a 33 M⊕ (0.1 MJupiter)
protoplanet in a disc with Σp = 75 g cm−2, and interstellar grain opacity.
Overlaid is a fit of the form M = atb + c (dashed line), where the fit is made
over the range 10 − 50 orbits (shaded) and matches the measured growth
over the entire evolutionary period. The upper line is the accretion history
for a 33 M⊕ protoplanet, in a disc of Σp = 750 g cm−2, with 0.1 per cent
IGO (model J), which by virtue of these conditions accretes much more
rapidly. A fit to this curve is also shown. The departure of the evolution-
ary model from the fitted curve is as discussed in Ayliffe & Bate (2009a).
The accretion rate initially declines as gas accretion builds a more optically
thick envelope, and trapped heat prevents the envelope contracting, slowing
its further accretion. However, when a sufficient mass is accreted, gravity
comes to dominate the growth process, resulting in an accelerating accre-
tion rate.
riod during which it is almost constant. However, such an initially
declining accretion rate was also seen in the semi-analytic models
of Papaloizou & Nelson (2005), where the planetary accretion rates
initial decreased due to the significant liberation of binding energy
as mass fell into the planet's potential. This energy release heats the
forming envelope, increasing the pressure, and reducing the rate
at which mass can be added. Eventually, however, in all models,
once a critical mass is reached, the energy release within the enve-
lope is insufficient to further retard the accretion, which then accel-
erates; this marks the transition from the thermally-dominated to
gravitationally-dominated accretion regime.
In Ayliffe & Bate (2009a), we studied the initially declin-
ing accretion rate in one case which we followed for 160 orbits,
but which did not reach the gravitationally-dominated regime. We
found that the slowing envelope growth could be represented very
well with an analytic fit of the form M = atb + c where b = 0.40.
In our new models, Model J is similar to the model from Ayliffe &
Bate (2009a), but with a much lower opacity and a higher disc sur-
face density. This produces higher accretion rates and, thus, allows
us to reach the gravitationally dominated growth regime. We refit
the old model using only orbits 10 − 50, and obtain essentially the
same fit we obtained in Ayliffe & Bate (2009a) using all 160 orbits;
© 0000 RAS, MNRAS 000, 000 -- 000
6
B.A. Ayliffe & M.R. Bate
Figure 4. The accretion history of the gaseous envelope in model J, where
the accreted mass is that contained within the self-consistently calculated
Hill radius (solid line). Also shown is the mass evolution with time within
radii of 0.3 (dashed), 0.1 (dot-dash), and 0.03 RH (dots-dash). The envelope
undergoes a hydrodynamical collapse after around 190 orbits when its mass
is ≈ 0.375 MJupiter, with the resulting shock pushing material out of the Hill
radius, whilst the remaining mass becomes more centrally condensed. This
phase is shown in the inset panel, which illustrates the central condensation
by the increasing mass within small radii as the overall mass falls. Follow-
ing the collapse, accretion resumes, replacing the mass lost due to the shock
propagation.
grain opacity. The mass evolution of this protoplanet is shown in
Fig. 4. As in all cases, the accretion rate initially drops away over
the first ∼ 50 orbits as the model settles, before beginning to in-
crease. The acceleration of accretion becomes more marked at the
crossover mass of ≈ 0.1 MJupiter, at around 100 orbits and continues
during the runaway growth phase. A change occurs at around 190
orbits, when the accreted mass has reached ≈ 0.375 MJupiter. At this
point the mass of the envelope is such that it overcomes the pres-
sure gradient which hitherto had supported it, and only by centrally
condensing can a new hydrostatic equilibrium be established. To
achieve this the envelope rapidly collapses, in the process trigger-
ing a shock wave which propagates outwards from the very dense
core that forms. This shock pushes material outwards, removing
a small fraction of the material from within the Hill radius very
rapidly at ≈ 190.5 orbits. The drop in the mass within 1, 0.3, and
0.1 RH can be seen clearly in the inset panel of Fig. 4. Most interest-
ingly, the reduction of the mass within 0.1 RH indicates that there
is a small loss of bound material as this radius falls well within the
expected limit of a protoplanetary envelope. The magnitude of the
mass loss from each volume is ∼ 1 per cent of the mass therein.
Following a brief increase (≈ 190.6 − 191 orbits), the mass within
RH continues to decline over the course of a further orbit before
it begins to grow once again. However, at the same time the mass
within 0.1 RH increases rapidly, showing that the available material
is quickly condensing around the core. Beyond ≈ 192 orbits the
accretion rate measured at the Hill radius resumes its pre-collapse
rate of 4 × 10−4 MJupiter year−1. Moreover, this rate appears to be
consistent at nearly all radii, indicating that gas is falling directly
onto the new denser envelope, rather than becoming suspended in
a more extended structure as was the case prior to collapse where
the accretion rates were different at different radii.
The process of envelope collapse is presented in Fig. 5, the
left hand panels of which shows cross-sections in density in the
Z-X plane through the centre of the planet, from a time preceding
the collapse in the uppermost panel, and at various stages during the
collapse in subsequent panels. In the first panel the density contours
are near spherical at small radii, elliptical at 0.5 RH, and pinched in
towards the planet's poles at large radii. The second panel illustrates
the strong shock propagating outwards from near the planet's core,
which at this point has not altered the gas structure near the Hill
radius (marked with a dotted line). However, it is possible to see
that the collapse has been greater at the poles, deforming the previ-
ously elliptical contours. The third panel demonstrates the contin-
ued propagation of the shock, which is elongated along the vertical
axis due to the lower density of material above and below the plane
of the disc, which allows the shock to propagate more rapidly in this
direction. Contours which had been pinched in at the planet's poles
are forced outwards as the shock passes, but as can be seen, inside
they are already resuming their pinched structure. The bottom panel
illustrates the resettled state of the protoplanet and its surround-
ings. The pinching at the poles, which in the top panel was found
beyond 0.5 RH, now extends down to ≈ 0.1 RH, and the surround-
ing medium now forms a circumplanetary disc, having previously
existed as an ellipsoidal envelope. The central density of the proto-
planet, that is the gas density near the core, was ≈ 5 × 10−4 g cm−3
in the top panel, and has increased to ≈ 5.5 × 10−3 g cm−3
in the
structure shown in the bottom panel following the collapse.
Returning to the formation of a circumplanetary disc, Fig. 6 il-
lustrates the midplane and vertical density structures about the pro-
toplanet, emphasising the altered state of the envelope. The diver-
gence from a spherically-symmetric density distribution, defined
here as a difference of greater than 10 per cent between the mid-
plane and vertical distributions, occurs at a radius 5 times smaller in
the post collapse state than in the pre-collapse state. This illustrates
that the envelope undergoes a more significant change in its struc-
ture vertically, than it does in the plane of the disc, and is similar
to the results seen in Ayliffe & Bate (2009a). This was investigated
further in Ayliffe & Bate (2009b) where it was found that circum-
planetary discs formed around massive protoplanets, and that these
discs tended to be thick, with dimensionless scale heights generally
larger than 0.2. We measure the circumplanetary disc scale height
by taking radial bins within the region with r < RH/3 (measured
from the protoplanet). In each radial bin, we take the SPH particle
densities (from all azimuthal angles) and fit Gaussian profiles to the
resulting vertical density distributions. A Levenberg -- Marquardt al-
gorithm is used to perform the fit, allowing the scale height to vary.
This gives a measure of disc scale height versus radius, which in
this case yields values for H/r that increase from 0.4 to 0.5 over
the radial range of 0.05−0.33 RH over which the disc extends. This
radial extent of the circumplanetary disc can be seen in Fig. 7 in
which the disc edge is taken to be at the location where the peak of
the specific angular momentum occurs. Beyond this radius, the spe-
cific angular momentum decreases with radius since, in the frame
rotating with the protoplanet, the material in the circumstellar disc
is counter-rotating relative to the gas captured by the protoplanet.
Over the radial range 0.07−0.3 RH the specific angular momentum
is on average 0.65 of the Keplerian values (marked by the dashed
line). The displacement is due to the pressure support within the
thick circumplanetary disc, and the degree of this displacement can
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
7
Figure 5. Cross-section plots illustrating the envelope collapse about a 33 M⊕ (≈ 0.1MJupiter) core. The Hill radius is marked as a dotted line. The accreted
mass (gas within RH) is ≈ 0.375 MJupiter prior to the collapse, reducing by ≈ 1.3 per cent as the shock pushes material away from the core. The x and z axes
are given in units of rp. Left panels: Cross sections in density illustrate the envelopes collapse, the shock propagation, and the formation of a circumplanetary
disc. The central density increases by an order of magnitude from the top panel to the last. Right panels: Temperature cross-sections at equivalent times to the
density panels. The peak temperature increases by 2100K to 6800K from the first to the last panel. The protoplanetary disc scaleheight is ≈ 0.05 rp, thus there
is little material obstructing the vertical propagation of the shock front. As a result, the shock propagates more easily in the vertical direction than through the
denser midplane, yielding the non-spherical propagation most clearly shown in the third righthand panel.
© 0000 RAS, MNRAS 000, 000 -- 000
z-0.0500.05z-0.0500.05z-0.0500.05zx0.9511.05-0.0500.05-12-10-8-6-4log density [g/cm3]z-0.0500.05z-0.0500.05z-0.0500.05zx0.9511.05-0.0500.0522.533.5log Tgas8
B.A. Ayliffe & M.R. Bate
Figure 6. The density distribution along the x-axis (solid lines) and along
the z-axis (dashed lines) through the protoplanet before (top panel) and af-
ter (bottom panel) collapse. The plus symbols mark the radius and density
at which the midplane and vertical density distributions differ by more than
10 per cent. The vertical dotted line marks the protoplanet's Hill radius. The
density distribution is modified significantly in both the midplane and ver-
tically following collapse, but the point at which these distributions diverge
moves inwards by a factor of 5 in radius.
be used to approximately calculate the disc scale height as another
check on the values measured above. Using the ratio of the specific
angular momentum ( j) to the Keplerian value ( jk) we obtain a scale
height of 0.55. This is obtained using equation 3 (see appendices B
& C of Laibe et al. 2012), where we have used values of 1/2 and
7/10 for the surface density and temperature exponents (p and q)
for the circumplanetary disc, typical values from Ayliffe & Bate
(2009); the calculated scale height is not enormously sensitive to
variations in these values within a reasonable range.
(cid:115)
H
r
=
2(1 − j/ jk)
p + q/2 + 3/2
(3)
This resulting scale height is somewhat larger than that which we
directly measured, but has been calculated assuming a vertically
isothermal disc and a lack of self-gravity, and is thus only approxi-
mate. The specific angular momentum distribution further supports
our assertion that a circumplanetary disc has formed as a result of
the envelope collapse.
The right hand panels of Fig. 5 shows the temperature struc-
ture in a Z-X slice at equivalent times to those shown in the density
panels to the left, allowing us to see the temperature evolution dur-
ing the collapse and shock propagation. A hot front associated with
the shock can be seen expanding away from the core with time,
and the vertical elongation of the shock can be clearly seen in the
third panel. Near the protoplanet's core, the peak temperature has
increased by 2100K to more than 6800K in the post-collapse state
shown in the final panel, and continues to increase for the remain-
der of the calculation when accretion has recommenced.
Fig. 8 shows the spherically averaged radial density and tem-
perature profiles running outwards from the protoplanetary core at
a number of points in the calculation. The change between 50 and
Figure 7. The specific angular momentum of the gas surrounding the proto-
planet that comprises the circumplanetary disc. The vertical dot-dashed line
marks RH/3, the analytically expected edge of the disc, and this matches
the measured turnover very well. The dashed line marks the Keplerian or-
bital velocity based on the mass within the associated radius. Pressure sup-
port within the disc means that a sub-Keplerian orbital velocity is expected,
though with a similar gradient if the disc is rotating about the planet. This
gradient matches reasonably between 0.07−0.3 RH.
150 orbits (the dashed lines) is relatively small, increasing as would
be expected for a growing protoplanet. Arriving at the pre-collapse
state (thin solid line) at around 190 orbits, at which point the accre-
tion rate is at its maximum, the density and temperature maxima
have increased more over 40 orbits than in the preceding 100 or-
bits. Moreover, the temperature structure shows a marked change
in its form. However the most significant changes occur during the
collapse and 5 further orbits of evolution (thick solid line). The en-
velope's collapse occurs very rapidly, and only stops when the new
structure of the envelope is able to reestablish hydrostatic equilib-
rium. For this to occur the density structure changes to deliver a
much steeper gradient away from the planet's solid surface. This
leads to a much steeper pressure gradient in this region, as shown
in the bottom panel of Fig. 8, eventually satisfying the requirement
∇P = −ρ∇φ, where φ = GM(r)/r. As such, the envelope is able
to resume steady accretion as the structure is able to bear the in-
creasing weight; as mentioned previously, accretion resumes at the
pre-collapse rate.
Fig. 9 depicts the changing distribution of mass in the inner
envelope from its pre-collapse state, to its post-collapse state. The
total time span shown is 229 days, the time between pre and post
collapse within the inner envelope, with these states marked using
dashed lines. However, the most significant changes occur over less
than 22 days at this scale, and this period is broken down in Fig. 9
into 4.3 day increments which are marked with solid lines. The Hill
radius just prior to the collapse is equal to 0.054 rp, which taking
Lissauer et al. (2009)'s estimate of a ∼ 0.25 RH envelope radius,
gives a size of 0.0135 rp. This radius is in reasonable agreement
with the region over which the mass is significantly redistributed
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
9
Figure 9. Cumulative mass distribution calculated outwards from the core
of Model J, where Macc = 0.375 MJupiter. The mass distribution is shown
over times ranging from just before the collapse, to after the structure sta-
bilises. The collapse proceeds very rapidly, as shown by the solid lines with
cover a period of less than 22 days from first (lowest) to last (highest).
Figure 8. Spherically averaged density (upper panel), temperature (middle
panel), and pressure (lower panel) distributions about the protoplanet. The
dashed lines are at times of 50, 100, and 150 orbits, their order ascending up
the left hand axis. The solid lines show the pre-collapse state (thinner line),
and the post-collapse state (thicker line). The pre-collapse state is equivalent
to that in Fig. 9, whilst the post-collapse state is instead taken at the very end
of the calculation in this case, when accretion has resumed its pre-collapse
rate.
during the collapse, which can be seen in Fig. 9 to be 0.01 rp (≈
0.2 RH).
Collapse of the protoplanetary envelope leads to a substantial
increase in the temperature near the protoplanetary core, as dis-
cussed above. A result of this temperature rise is an increase in the
dissociation fraction of molecular hydrogen about the core, and an
© 0000 RAS, MNRAS 000, 000 -- 000
Figure 10. Fraction of atomic hydrogen versus radius within the inner re-
gion of the protoplanetary envelope before it collapses (dashed line), and
after it collapses (solid line). The higher temperatures that develop within
the deep envelope when it collapses lead to a higher dissociation fraction of
molecular hydrogen, whilst this process of dissociation will absorb energy,
reducing the maximum temperature that is achieved. Prior to collapse the
fraction of atomic hydrogen peaks at 0.33, whilst in the immediate after-
math it is as high as 0.48.
10
B.A. Ayliffe & M.R. Bate
enlargement of the region within which hydrogen is substantially
dissociated; this can be seen in Fig. 10. It is not the dissociation that
triggers the collapse of the envelope, despite the process acting as
an energy sink; the capacity of dissociation to absorb energy does
lead to a lower final temperature within the envelope than might
otherwise have been achieved.
3.3 Accretion flow
Whilst we find that a protoplanet envelope increases in mass
throughout its evolution, excepting a brief period following a dy-
namical collapse (seen at radii of 1, 0.3, and 0.1 RH in Fig. 4), it is
not obvious that this accretion is a spherically symmetric process.
Machida et al. (2008) and Tanigawa et al. (2012) have performed
three-dimensional calculations of protoplanet growth, and find that
gas flows outwards along the midplane from a growing protoplanet,
such that accreted material must be delivered vertically. In Ayliffe
& Bate (2009b) we found that mass predominantly entered the Hill
sphere along the midplane, but this analysis was simplistic in that
it only considered inflow, failing to consider the possibility of out-
flow, and we worked solely with the azimuthally integrated values.
Here we make a more thorough assessment of the mass flow, and
are able to look at this flow in a large extended envelope that has not
undergone collapse, as well as in the dense envelope formed sub-
sequent to such a collapse. It is the latter case which most closely
resembles the principle model of Tanigawa et al. (2012) for a high
mass protoplanet.
Before and after the envelope collapse, we see material flow-
ing in and out at the Hill radius in an alternating pattern, as can
be seen in the first panels of Figs. 11 & 12, where negative values
correspond to inflow, and the solid contour line denotes a flux of
zero. This flow is easily explained as the passage of gas passing
the planet and being deflected by the spiral shocks that form due to
the gravitational perturbation provided by the protoplanet. Another
contribution comes from gas following horseshoe orbits which also
enter and leave the Hill sphere at broadly similar longitudes about
the planet. Fig. 13 illustrates the vector field that results in the mass
flow observed at the Hill radius, which is marked in this figure with
a dashed line. The data presented in Figs. 11 & 12 was constructed
by considering the flow over a period of 4 orbits preceding and fol-
lowing the collapse respectively. The stability of the gas flow about
the protoplanet over these periods leads to the regular pattern that
is seen at the Hill radius in both figures. In the pre-collapse case the
pattern persists down to 0.3 RH as the second panel of Fig. 11 illus-
trates. At smaller radii the flow is inwards across the spherical shell,
with a larger flux at smaller radii as is expected for a consistent
mass flowing across a decreased surface area. In the pre-collapse
state the mass flow at small radii does not appear to possess any
latitudinal dependence, rather it flows in almost spherically sym-
metrically. Note that the figures are noisey at the poles as a result
of the spherical polar grid used to calculate the flux, which leads to
very small bins at high latitudes.
In the post-collapse case, shown in Fig. 12, the mass flux at
the Hill sphere is little changed from the pre collapse case, and in
both cases it is reminiscent of the structures seen in Fig. 5 of Tani-
gawa et al. (2012); note that we are plotting a time average flux,
whilst Tanigawa et al. plot an instantaneous flux (ρvr). Unlike Tani-
gawa et al. (2012), who see these alternating structures persisting
down to very small radii, our model has already lost any sign of
the in-out flow pattern at a radius of 0.3 RH. Instead, at 0.3 RH we
find an outflow along the circumplanetary disc midplane at every
longitude, though this outflow shows a slightly alternating magni-
tude, with more significant outflow at longitudes of ∼ 20 and 200
degrees. This outflow originates at a radius of ≈ 0.17 RH, which
is around half the radius of the circumplanetary disc, inside of this
radius the flow is inwards. Meanwhile at 0.3 RH, mass is flowing
inwards at higher latitudes. The relative fluxes are such that the net
flux at each radius is negative, enabling the protoplanet to continue
to grow. This growth is corroborated by Fig. 4, which shows the
mass evolution of Model J within the 4 radii considered here, and
indicates that the mass consistently increases within 0.3 RH (dashed
line) once the post-collapse state is achieved. Following the enve-
lope's collapse, there is more discernible structure to the mass flow
at the smaller radii. At 0.1 RH the inflow along the circumplanetary
disc midplane is not as significant as the flow at high latitudes, as
is shown in the third panel of Fig. 12. This panel marks a clear di-
vide in the motion of gas at and around a radius of 0.1 RH, where
the flow is only found to be inwards, signalling that material within
this radius is truly bound to the protoplanet. This is of particular
interest because of the gas flow found at 0.03 RH, and shown in
the final panel of Fig. 12. At this small radius there are signifi-
cant flows of material, pushing outwards at intermediate latitudes
(∼ 45−70 degrees), and pouring inwards again along the midplane.
This is indicative of significant circulation of the bound material
below 0.1 RH, and will be discussed further in Section 4.2.
4 DISCUSSION
4.1 Hydrodynamic collapse
From the earliest suggestion of Perri & Cameron (1974) it has been
thought that a giant planet might form through the hydrodynamic
collapse of a gaseous envelope onto a solid core which caused it
to assemble. This was followed by numerous models that effec-
tively sought for hydrostatic solutions to various combinations of
properties to establish when such a collapse might occur (Mizuno
et al. 1978; Mizuno 1980; Sasaki 1989). The first models that at-
tempted to model giant planet growth from the initial core forma-
tion, through to the envelope growth were performed by Boden-
heimer & Pollack (1986). These models revealed that a protoplan-
etary envelope would gradually contract as the planet grew, leading
to a quasi-static contraction beyond previously calculated values
for the critical mass, as long as the envelope did not effectively de-
tach from the protoplanetary disc (that is there was a sufficiently
rapid supply of material from the latter to the former). Under these
conditions, their was no evidence to suggest that the hydrostatic
balance should reach some limit beyond which a collapse was in-
evitable, and later semi-analytic models originating from these ear-
lier works, such as Pollack et al. (1996), Hubickyj et al. (2005), and
Lissauer et al. (2009), suggest no need for a dynamic collapse. Our
Model J follows a pattern of stable growth for the vast majority of
its history, though not evidently undergoing any significant contrac-
tion, and resumes this pattern of growth subsequent to its envelope
collapse.
The models presented in this article are performed using a
three-dimensional hydrodynamics code that include self-gravity,
and radiative transfer, but which omits the core formation phase,
and the deposition of energy due to planetesimal accretion that are
included in the semi-analytic works discussed. However, at the time
of interest around the envelope collapse, the energy release is ut-
terly dominated by the contraction of the gaseous envelope, such
that solids accretion energy may be regarded as negligible. The
metallicity of the envelope might be significantly modified by the
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
11
Figure 11. Mass flux through shells of various radii (marked in panels) surrounding the protoplanet core in Model J prior to its envelope collapse. At one Hill
radius, the flow takes on a form that is similar to that seen in Fig. 5 of Tanigawa et al. (2012), and reflects the combined effect of horseshoe orbits and bent
flow lines passing the protoplanet. Material both enters and leaves the sphere predominantly at the midplane where densities are highest; the marked contour
line denotes a flux of zero such that the area within the contour marks the outflow. For ease of comparison with Tanigawa et al. (2012) the solar and anti-solar
points are at longitudes of 0 and 180 degrees respectively, and inflowing material is shown by a negative flux. At the smallest radii the flow is largely inwards
across the shell, with an average flux of −1 × 10−2 code units (−5.5 × 10−5 g cm−2 s−1).
© 0000 RAS, MNRAS 000, 000 -- 000
12
B.A. Ayliffe & M.R. Bate
Figure 12. Identical to Fig. 11 except illustrating the gas flow after Model J undergoes its envelope collapse. The flow at one Hill radius is very similar to the
pre-collapse case, however at radii between 0.1− 0.3 RH the outflow is concentrated along the midplane, whilst inflow occurs at higher latitudes, reflecting the
vertical accretion seen by Machida et al. (2008). At the smallest radii, there is evidence of a gas turn over, where material is flowing in along the midplane and
at the poles, and out in a range of moderate to high latitudes (see Section 4.2).
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
13
Figure 13. Velocity vectors in the plane of the disc, illustrating the gas flow
about the Hill radius (marked with a dashed line) that leads to the alternating
pattern of in and out flow seen in the first panels of the mass flux plots
shown in Figs. 11 & 12. This vector field is plotted for a time preceding the
envelope collapse, but a very similar field exists after the shock associated
with the collapse has passed out of the region.
ablation and evaporation of grains that have been accreted over the
planets history, and we make no attempt to account for this. The
opacity in Model J is reduced by a fixed factor of 103, at the lowest
end of the range suggested by Movshovitz et al. (2010), who found
such opacities due to grain settling and coagulation in regions of
the envelope.
It is difficult to disentangle the causes and effects of the very
rapid collapse we find, for example the surge in temperature leads
to a higher fraction of dissociated hydrogen. However, as stated
in Section 3.2, it does not appear that the dissociation of molec-
ular hydrogen acts to trigger the collapse, as the fraction remains
steady in the preceding period. There is also no evidence of the
Kappa-mechanism acting within the protoplanetary atmosphere in
our model, as was found by Wuchterl (1991) to cause a dynamic
collapse. Wuchterl also found that this collapse led to a signifi-
cant ejection of material from the protoplanet, whilst we find only a
small drop of ≈ 1.3 per cent in the mass within the Hill radius, and
a rapid increase at and below 0.1 RH as the protoplanet structure
shifts to its new state.
It appears that our Model J protoplanet reaches the hydrostatic
limit for its formative structure, and that the internal pressure gra-
dient can no longer accommodate the addition of mass by a small
adjustment. This is illustrated in Fig. 14 which shows the ratio
of the pressure force to the gravitational force against radius at a
number of stages of the envelope collapse. From an initial state,
in which the protoplanetary atmosphere is in hydrostatic equilib-
rium out to a radius of 0.0067 rp (0.12 RH), the atmosphere rapidly
begins to restructure, pushing the hydrostatic region down to a ra-
dius of ≈ 0.0013 rp (0.025 RH). This initial stage leaves the form
of the graph otherwise relatively unchanged, but as the central con-
centration of mass continues, a shock begins to form as the pres-
sure near the core surges. The maximum pressure within 0.001 rp
has increased by an order of magnitude between the first panel and
the third, leading to a somewhat steeper gradient over this region.
However, at this point in time the gradient between 0.001−0.002 rp
steepens much more rapidly, forming a shock front, and this front
marks the new radial limit of hydrostatic equilibrium as can be seen
in the third panel. The subsequent panel shows the shock prop-
© 0000 RAS, MNRAS 000, 000 -- 000
Figure 14. Each panel shows the ratio of the pressure force to the gravita-
tional force in the envelope and beyond, with the first and last panels corre-
sponding in time to the pre and post-collapse cumulative mass distributions
shown in Fig. 9. A ratio of one indicates that the material is in hydrostatic
equilibrium, which before the envelope collapse applies to a region out to
0.0067 rp (0.12 RH, top panel, marked with vertical dotted line), but which
post-collapse reaches out to just 0.0009 rp (0.016 RH). By the final panel
the shock wave has cleared the inner 0.01 rp, leaving the environment in-
ternal to this in a new hydrostatic equilibrium, whilst its mass continues to
increase, as shown in Fig. 4.
14
B.A. Ayliffe & M.R. Bate
agating outwards, whilst the hydrostatic core shrinks a little. By
the final panel the inner envelope has resettled and the structure
has stabilised, whilst the shock's propagation continues outwards,
eventually moving beyond the limits of the modelled region.
It is possible that this envelope collapse only occurs due to the
high accretion rates achieved due to our selected disc conditions.
The low opacity assumed promotes very rapid planet growth, and
it may be this rapidity that prevents the envelope from adjusting
its structure more gradually to accommodate the increasing mass.
As such, it may be that the hydrodynamical collapse of a plan-
etary atmosphere can only occur if that planet if accreting very
rapidly. Further models will be required to determine whether or
not this is the case. We note however that the accretion rate mea-
sured in Model J, both just before and after the collapse, is still
not as rapid as would be found using a locally-isothermal equa-
tion of state, despite the large reduction in opacity. Lissauer et al.
(2009) present accretion rates in their fig. 3, where these rates were
obtained by D'Angelo et al. (2003b) using three-dimensional hy-
drodynamical models with a locally-isothermal equation of state.
Applying our disc conditions to their results yields an accretion
rate of 8 × 10−4 MJupiter year−1, which is twice the rate measured in
our radiation hydrodynamics models prior to collapse.
At this juncture we note that the results given in Lissauer
et al. (2009) show a viscosity dependence, where higher viscosi-
ties lead to more rapid accretion. In our calculations viscosity is
not a constant, but is proportional to the spatial resolution of the
SPH method. Thus, the viscosity is lower in regions of higher den-
sity, and these differences mean the above comparison is only ap-
proximate. Further, we note that in the absence of a protoplanet,
the unperturbed circumstellar disc in our models has a viscosity of
α ≈ 4 × 10−3, consistent with the fixed viscosity global models of
Bate et al. (2003) that are used to inject gas at the boundaries of
the disc section. It is these boundaries that determine the rate at
which gas is supplied to the disc section in these local models. As
such, once the disc is perturbed, the spatially varying viscosity of
the SPH calculations leads to to an inconsistency with the global
models, and so the boundaries. A further caveat arising from the
boundary implementation is that the injected material comprises
gas on both circulating and librating orbits (Lubow et al. 1999), or-
bits that are modified as the gas passes through the modelled disc
section. However, these modifications are lost when the gas leaves
the section, and new gas is injected without these modifications,
leading to a further inconsistency.
4.2 Atmospheric turn over
As briefly mentioned in Section 3.3 in reference to the third panel
of Fig. 12, there appears to be significant motion of the bound gas
besides rotation in the plane of the disc. An apparent rolling mo-
tion is indicated by the flow through the surface at 0.03 RH, the
gas unable to escape, but flowing out and in within the bound at-
mosphere. These flows appear to be convection cells in the deepest
parts of the planetary envelope where the medium is extremely op-
tically thick, and thus unable to cool readily by radiation. Fig. 15
shows a region out to 0.07 RH (3.7 × 10−3 rp), at which radius the
temperature is 675K, increasing to a maximum of 7400K near the
core, demonstrating a significant thermal gradient against radius.
Bodenheimer & Pollack (1986) found strata of convection within
protoplanetary envelopes in their one-dimensional models, with a
dense inner convection region that contained some 95 per cent of
the envelope mass, and depending on the age and core size, mul-
tiple higher level convection zones were also found to form. How-
Figure 15. A slice through Model J after the envelope has collapsed, show-
ing the mean velocity vector field over the course of 4 orbits which corre-
spond to those used to produce the mass flux plots in Fig. 12. The outer
circle corresponds to the spherical surface of Fig. 12 at 0.03 RH, whilst the
inner filled circle illustrates the radius of the planetary core. This slice in
the z-rcyl plane rotates about the core at the same average angular velocity
as the gas, which exhibits solid-body rotation within 0.04 RH, so that the
vector field is seen in the rotating frame of the gas. The bold sections of the
outer circle cover the latitudes of 45 − 70 degrees at which out flow was
seen in the final panel of Fig. 12.
ever, it is likely that the convection in fully 3-D models that include
rotation, is very different.
Fig. 15 illustrates the vector field in a slice through the proto-
planet, averaged over 4 orbits that correspond to those used to pro-
duce the mass flux rendering of Fig. 12. Within ∼ 0.04 RH the proto-
planet rotates as a solid body, with a rate of 4.65×10−7 rad s−1 (giv-
ing a day equivalent to ≈ 160 Earth days). The plane to which the
velocity vectors have been mapped was rotated at this rate, such that
the inner region of the atmosphere remained consistently aligned
with the plane over the time of averaging. There are four distinct
regions of turn over in which gas flows outwards at mid-latitudes
before falling inwards again along the disc midplane. These looping
flows are subsonic, with mean velocities of less than 3 m s−1 and a
maximum of less than 8 m s−1 in a region where the sound speed is
greater than 6 km s−1. A time average to a fixed plane, or a spatial
average rotating about the z-axis at a single moment in time both
reveal a similar structure, suggesting that the general form of these
structures is persistent. The bold sections of the 0.03 RH circle are
marked between 45 − 70 degrees, the latitudes where outflow was
seen in the fourth panel of Fig. 12. As might be expected, the veloc-
ity vectors across these bold segments indicate outflow, illustrating
that these looping flows are responsible for the features in the ren-
dered plot. The 0.03 RH region marked by the outer circle contains
≈ 97 per cent of the total protoplanetary mass, that is the mass mea-
sured out to the Hill radius; prior to collapse the same region had
contained just under 40 per cent of the mass within the Hill radius.
In this dense environment, possessing a significant radial tempera-
ture gradient, the 4 distinct cells revealed in the velocity field might
well be indications of convection.
© 0000 RAS, MNRAS 000, 000 -- 000
Hydrodynamic collapse models
15
Figure 16. Longitudinally (azimuthally) integrated mass flux for Model J at the same four radii considered in Figs. 11 & 12. The solid lines represent the
pre-collapse case (Fig. 11) and the dashed lines the post-collapse case (Fig. 12). In the former, it is notable that the only instance in which an outward flux
can be seen is for the largest radii, where this is easily understood by consideration of the large scale flows (spiral shocks, horseshoe regions, and vertical
accretion) surrounding the planet, and persists in a similar form post-collapse. Post-collapse there is more structure visible in the flow at small radii. At 0.3 RH
there is a midplane outflow that corresponds to that visible the corresponding panel of Fig. 12, whilst at 0.1 RH all material is flowing inwards, though at the
midplane the flux is small relative to other latitudes. Finally, at 0.03 RH there is evidence of significant turnover in the bound envelope post-collapse, possibly
indicative of convection.
4.3 Gas accretion
Tanigawa et al. (2012) have recently examined gas flow onto and
within circumplanetary discs. They found, in agreement with the
work of Machida et al. (2008), that gas flowed onto a circumplan-
etary disc predominantly in the vertical direction. Moreover, their
works suggests that material is flowing out along the midplane of
a circumplanetary disc. Once the envelope collapses in Model J of
this work, and a circumplanetary disc forms, we also find that ma-
terial is flowing outwards along the midplane beyond a radius of
0.17 RH, which is at around half the outer radius of the circum-
planetary disc (Quillen & Trilling 1998; Ayliffe & Bate 2009b;
Martin & Lubow 2011). Fig. 16, which is equivalent to Fig. 6 in
Tanigawa et al. (2012), illustrates the longitudinally (equivalently,
azimuthally) integrated mass flux for the pre (solid line) and post
(dashed line) collapse flows shown in Figs. 11 & 12. Before the
envelope collapses, at 0.3 RH the outflow seen in two places at the
midplane is counterbalanced by the associated midplane inflows
and the inflow at higher latitudes, yielding a net inflow, as shown by
the solid line in the second panel of Fig. 16. However, post collapse
the consistent midplane outflow seen for this radius at all longitudes
results in a peak of outflowing material in a region ±30 degrees lat-
itude.
© 0000 RAS, MNRAS 000, 000 -- 000
About the planet, at one Hill radius the flow is dominated by
the streams of material passing in and out of the region due to the
form of their circumstellar orbits (see Fig. 13). As such, it is un-
surprising that the fluxes we obtain, and those of Tanigawa et al.
(2012) scaled to our model, are similar at this radius; note we make
the following comparisons using our post-collapse case which bet-
ter resembles Tanigawa et al.'s models. At 1 RH they obtain a peak
outflow along the midplane of 1.8 × 10−7 g cm−2 s−1, whilst we
obtain a value of 2.5 × 10−7 g cm−2 s−1. The form is also similar,
with wings of inflow at higher latitudes of similar magnitude; at
the highest latitudes, our normalisation by area leads to a non-zero
inflow. However, at smaller radii the mass fluxes we find are con-
siderably larger than at the Hill radius, which differs substantially
from Tanigawa et al.'s results, where the peak flux at all radii dif-
fer by less than an a factor of 4. At 0.3 RH the peak outward flux
has grown to 2.2 × 10−6 g cm−2 s−1, and at 0.1 RH the outflow has
ceased, but the inflow flux at high latitudes has increased by an-
other factor of around 5. The final panel of Fig. 16 reveals the mass
flow resulting from the formation of convection cells in the deep
atmosphere post collapse.
16
B.A. Ayliffe & M.R. Bate
5 SUMMARY
We have performed three-dimensional self-gravitating radiation hy-
drodynamical models of planet growth that, subsequent to an ex-
tended period of growth, result in a dynamical collapse. A series
of models were performed using different core masses, core radii,
and opacities, that extend the range of accreted mass achieved in
Ayliffe & Bate (2009a). We present the first results from a three-
dimensional hydrodynamical model of planet growth by core accre-
tion that has been found to produce a hydrodynamic collapse. The
result of this collapse is a very centrally-condensed protoplanet,
surrounded by a circumplanetary disc, that continues to accrete.
The inner reaches of the envelope have undergone significant dis-
sociation of molecular hydrogen, and appear to possess convection-
like cells of gas turn over, whilst the inflow of new gas occurs near
vertically at high latitudes.
The circumplanetary disc, with radius ≈ RH/3 and dimension-
less scaleheight of 0.4 − 0.5, exhibits a reversal in the direction of
mass flow along the midplane at around 50 per cent of its radius;
that is to say there is inflow only within the inner 0.17 RH. The
degree of central condensation in the post-collapse state leads the
model to show good agreement with previous calculations consid-
ering mass flow that have presumed a pre-existing high mass core
of relatively small size (Tanigawa et al. 2012). Conversely, before
the collapse, the extended protoplanetary envelope exhibits a more
spherically symmetric inflow of material.
To achieve rapid growth in these models we have adopted very
favourable conditions, particularly a low opacity of just 0.1 per cent
of the interstellar grain opacity. It may be that the hydrodynamic
collapse found in this work is a result of the very rapid growth of
the envelope, promoted by these disc conditions, though this can-
not be said definitively without performing further models in less
favourable discs.
ACKNOWLEDGMENTS
We would like to thank the anonymous referee, whose comments
helped us to clarify our results. We would also like to thank Pe-
ter Bodenheimer for his comments on the manuscript. The calcu-
lations reported here were performed using the University of Ex-
eter Supercomputer. Several figures were created using SPLASH
(Price 2007), a visualisation tool for SPH that is publicly avail-
able at http://users.monash.edu.au/∼dprice/splash/. MRB is grate-
ful for the support of a Philip Leverhulme Prize and a EURYI
Award which also supported BAA. This work, conducted as part of
the award "The formation of stars and planets: Radiation hydrody-
namical and magnetohydrodynamical simulations" made under the
European Heads of Research Councils and European Science Foun-
dation EURYI (European Young Investigator) Awards scheme, was
supported by funds from the Participating Organisations of EURYI
and the EC Sixth Framework Programme.
REFERENCES
Alexander D. R., 1975, ApJS, 29, 363
Alibert Y., Mordasini C., Benz W., Winisdoerffer C., 2005, A&A,
434, 343
Ayliffe B. A., Bate M. R., 2009, in American Institute of Physics
Conference Series, Vol. 1158, American Institute of Physics
Conference Series, Usuda T., Tamura M., Ishii M., eds., pp. 219 --
221
Ayliffe B. A., Bate M. R., 2009a, MNRAS, 393, 49
Ayliffe B. A., Bate M. R., 2009b, MNRAS, 397, 657
Bate M., 1995, PhD thesis, PhD thesis, Univ. Cambridge, (1995)
Bate M. R., Lubow S. H., Ogilvie G. I., Miller K. A., 2003, MN-
RAS, 341, 213
Benz W., 1990, in Numerical Modelling of Nonlinear Stellar Pul-
sations Problems and Prospects, J. R. Buchler, ed., Kluwer Aca-
demic Publishers, Dordrecht, The Netherlands, pp. 269 -- +
Benz W., Cameron A. G. W., Press W. H., Bowers R. L., 1990,
ApJ, 348, 647
Bodenheimer P., Pollack J. B., 1986, Icarus, 67, 391
Boley A. C., Hartquist T. W., Durisen R. H., Michael S., 2007,
ApJ, 656, L89
Bryden G., Chen X., Lin D. N. C., Nelson R. P., Papaloizou
J. C. B., 1999, ApJ, 514, 344
D'Angelo G., Henning T., Kley W., 2003a, ApJ, 599, 548
D'Angelo G., Kley W., Henning T., 2003b, ApJ, 586, 540
Hubbard W. B., Macfarlane J. J., 1980, J. Geophys. Res., 85, 225
Hubickyj O., Bodenheimer P., Lissauer J. J., 2005, Icarus, 179,
415
Ikoma M., Nakazawa K., Emori H., 2000, ApJ, 537, 1013
Klahr H., Kley W., 2006, A&A, 445, 747
Laibe G., Gonzalez J.-F., Maddison S. T., 2012, A&A, 537, A61
Lissauer J. J., Hubickyj O., D'Angelo G., Bodenheimer P., 2009,
Icarus, 199, 338
Lubow S. H., Seibert M., Artymowicz P., 1999, ApJ, 526, 1001
Machida M. N., Kokubo E., Inutsuka S., Matsumoto T., 2008,
ApJ, 685, 1220
Martin R. G., Lubow S. H., 2011, MNRAS, 413, 1447
Mizuno H., 1980, Progress of Theoretical Physics, 64, 544
Mizuno H., Nakazawa K., Hayashi C., 1978, Progress of Theoret-
ical Physics, 60, 699
Monaghan J. J., 1992, ARA&A, 30, 543
Monaghan J. J., 2002, MNRAS, 335, 843
Monaghan J. J., Gingold R. A., 1983, Journal of Computational
Physics, 52, 374
Movshovitz N., Bodenheimer P., Podolak M., Lissauer J. J., 2010,
Icarus, 209, 616
Paardekooper S., Mellema G., 2008, A&A, 478, 245
Papaloizou J. C. B., Nelson R. P., 2005, A&A, 433, 247
Perri F., Cameron A. G. W., 1974, Icarus, 22, 416
Podolak M., 2003, Icarus, 165, 428
Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J.,
Podolak M., Greenzweig Y., 1996, Icarus, 124, 62
Pollack J. B., McKay C. P., Christofferson B. M., 1985, Icarus,
64, 471
Price D. J., 2007, Publications of the Astronomical Society of
Australia, 24, 159
Price D. J., Monaghan J. J., 2007, MNRAS, 374, 1347
Quillen A. C., Trilling D. E., 1998, ApJ, 508, 707
Sasaki S., 1989, A&A, 215, 177
Seager S., Kuchner M., Hier-Majumder C. A., Militzer B., 2007,
ApJ, 669, 1279
Slattery W. L., 1977, Icarus, 32, 58
Springel V., Hernquist L., 2002, MNRAS, 333, 649
Tajima N., Nakagawa Y., 1997, Icarus, 126, 282
Tanigawa T., Ohtsuki K., Machida M. N., 2012, ApJ, 747, 47
Whitehouse S. C., Bate M. R., 2006, MNRAS, 367, 32
Whitehouse S. C., Bate M. R., Monaghan J. J., 2005, MNRAS,
364, 1367
Wuchterl G., 1990, A&A, 238, 83
Wuchterl G., 1991, Icarus, 91, 53
© 0000 RAS, MNRAS 000, 000 -- 000
|
1702.03678 | 1 | 1702 | 2017-02-13T09:04:48 | Quantitative indexing and Tardigrade analysis of exoplanets | [
"astro-ph.EP"
] | Search of life elsewhere in the galaxy is very fascinating area for planetary scientists and astrobiologists. Earth Similarity Index (ESI) is defined as geometrical mean of four physical parameters (Such as radius, density, escape velocity and surface temperature), which is ranging from 1 (identical to Earth) to 0 (dissimilar to Earth). In this work, ESI is re-defined as six parameters by introducing the two new physical parameters like revolution and surface gravity and is called as New Earth Similarity Index (NESI). The main focus of this paper is to search Tardigrade water-life on exoplanets by varying the temperature parameter in NESI, which is called as Tardigrade Similarity Index (TSI), which is ranging from 1 (Tardigrade can survive) to 0 (Tardigrade Cannot survive). Here the NESI and TSI is cataloged and analyzed for almost 3370 confirmed exoplanets. | astro-ph.EP | astro-ph | Draft version October 13, 2018
Preprint typeset using LATEX style AASTeX6 v. 1.0
QUANTITATIVE INDEXING AND TARDIGRADE ANALYSIS OF EXOPLANETS
1Department of Physics, Jyoti Nivas College, Bengaluru-560095, Karnataka, India
J. M. Kashyap1
ABSTRACT
Search of life elsewhere in the galaxy is very fascinating area for planetary scientists and astrobiologists.
Earth Similarity Index (ESI) is defined as geometrical mean of four physical parameters (Such as
radius, density, escape velocity and surface temperature), which is ranging from 1 (identical to Earth)
to 0 (dissimilar to Earth). In this work, ESI is re-defined as six parameters by introducing the two new
physical parameters like revolution and surface gravity and is called as New Earth Similarity Index
(NESI). The main focus of this paper is to search Tardigrade water-life on exoplanets by varying
the temperature parameter in NESI, which is called as Tardigrade Similarity Index (TSI), which is
ranging from 1 (Tardigrade can survive) to 0 (Tardigrade Cannot survive). Here the NESI and TSI
is cataloged and analyzed for almost 3370 confirmed exoplanets.
Keywords: Revolution of exoplanets, Tardigrade Similarity Index, surface gravity of exoplanets, and
New Earth Similarity Index
1. INTRODUCTION
Exploring the unknown worlds outside our solar system (i.e., exoplanets) is the new era of the current research.
Presently with the huge flow of data from Planetary Habitability Laboratory PHL-HEC 1, maintained by university
of Puerto Rico, Arecibo. Indexing will be a main criteria to give a proper structure to these raw data from space
missions such as CoRoT and Kepler. Nearly half a decade ago, Schulze-Makuch et al.(2011) has defined Earth
Similarity Index (ESI) as a geometrical mean of four physical parameters (such as: radius, density, escape velocity and
surface temperature). In this paper, the New Earth Similarity Index (NESI) is re-defined as geometrical mean of six
physical parameters (namely: radius, density, escape velocity, surface temperature, revolution, surface gravity). Since
revolution is not directly available as the raw data, here the values are calculated for 3370 (as of September 2016)
confirmed exoplanets. We are always interested in life-forms, which can survive outside our planet and tardigrade is
the likely candidate (Jnsson et al. 2008). Tardigrade primarily known as moss piglets or water bears (Copley J. 1999),
and they can survive at different temperature scales (Examples: 151 deg C for few minutes (Horikawa, Daiki D. 2012),
−20 deg C for 30 years (Tsujimoto et al. 2015), −200 deg C for days (Horikawa, Daiki D. 2012), −272 deg C for few
minutes (Becquerel P. 1950) in lower temperature scale).
The structure of the paper is as follows: In section 2, the NESI has been introduced and analysis of 3370 confirmed
exoplanets are done, section 3 has the results of TSI, and section 4 gives the discussion and conclusion part of the
work.
In 2011, Schulze-Makuch et al., defined the Earth similarity index as
2. NEW EARTH SIMILARITY INDEX AND ITS ANALYSIS
7
1
0
2
b
e
F
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
8
7
6
3
0
.
2
0
7
1
:
v
i
X
r
a
(cid:20)
1 −(cid:12)(cid:12)(cid:12) x − x0
x + x0
(cid:12)(cid:12)(cid:12)(cid:21) wx
ESIx =
,
(1)
where x is the planetary property of the exoplanet, WX is the weight exponent and x0 is the reference to Earth in ESI.
In this paper we follow the above equation to calculate the NESI of individual planetary property, with the weight
exponents as defined below. The New Earth Similarity Index (NESI) is defined as the geometrical mean of six
[email protected]
1 http://phl.upr.edu/projects/habitable-exoplanets-catalog/data/database
2
J. M. Kashyap
physical parameters (such as: radius, density, escape velocity, surface gravity, revolution and surface temperature).
Mathematically it can be denoted as:
NESI = (NESIR × NESIρ × NESIT × NESIVe × NESIRe × NESIg)1/6
(2)
The weight exponents for upper and lower limits of parameters are calculated as (Schulze-Makuch et al.(2011)):
radius 0.5 to 1.9 EU, mass 0.1 to 10 EU, density 0.7 to 1.5 EU, surface temperature 273 to 323 K and escape velocity
0.4 to 1.4 EU. Similarly, we define the limits of gravity as 0.16 to 17 EU and revolution as 0.61 to 1.88 EU.
Gravity and Revolution:
The human centrifuge experiment clearly showed the untrained humans can tolerate 17 EU, with eye balls in (Brent
et al. 2012). The revolution is scaled on the basis of habitable zone of sun-like stars.
Table 1. NESI Parametric Table
Planetary Property
Ref. Value Weight Exponents
for NESI
for NESI
Mean Radius
Bulk Density
Escape Velocity
Revolution
Surface gravity
1EU
1EU
1EU
1EU
1EU
Surface Temperature
288K
0.57
1.07
0.70
0.70
0.13
5.58
Here, EU = Earth Units, where Earths radius is 6371 km, density is 5.51 g/cm3, escape velocity is 11.19 km/s,
Revolution is 365.25 days and surface gravity is 9.8 m/s2.
The NESI is further divided into Interior NESI and Surface NESI. The interior NESI is defined as the geometrical
mean of radius and density. Mathematically it takes the form:
NESII = (NESIR × NESIρ)1/2 ,
(3)
Similarly, the surface NESI is defined as the geometrical mean of surface temperature, escape velocity, revolution of
the planet and surface gravity. Empirically is of the form:
The global NESI takes the form:
NESIS = (NESIT × NESIVe × NESIRe × NESIg)1/4
NESI = (NESII × NESIS)1/2 ,
Table 2. A sample of calculated NESI
Names
Radius Density Temp E. Vel Rev
g
NESIS NESII NESI
(EU)
(EU)
Earth
Mars
Kepler-438 b
GJ 667C c
Kepler-296 e
1.00
0.53
1.12
1.54
1.48
1.00
0.73
0.90
1.05
1.03
(K)
288
240
312
280
303
(EU)
1.00
0.45
1.06
1.57
1.50
1.00
1.88
0.09
0.07
0.08
1.00
0.37
1.01
1.61
1.52
1.00
0.74
0.69
0.66
0.67
1.00
0.81
0.96
0.92
0.93
1.00
0.77
0.81
0.78
0.79
(4)
(5)
The entire analysis for 3370 exoplanets is cataloged in separate rocky and gas exoplanets files, which is available
at(Kashyap et al. 2017).
Quantitative indexing and Tardigrade analysis of exoplanets
3
Figure 1. Plot of interior NESI versus surface NESI. Blue dots are the giant planets, red dots are the rocky planets, and cyan
circles are the Solar System objects (Table 2). However, the limit for NESI is marked by a solid line which is ∼ 0.70.
The peaks in the plot are due to distribution phenomenon of similarity indices. This scatter plot denotes the
optimistic range of each and every gas giants and rocky exoplanets, with respect to Earth-like planets.
3. TARDIGRADE SIMILARITY INDEX
The Tardigrade Similarity Index (TSI) is defined similarly as NESI, with different weight exponent for surface
temperature. Mathematically it can be denoted as:
TSI = (TSIR × TSIρ × TSIT × TSIVe × TSIRe × TSIg)1/6
(6)
The weight exponent range for all the physical parameters are same as NESI, except the temperature parameter,
which ranges from 1.15 to 424.15 K for tardigrade to survive.
Table 3. TSI Parametric Table
Planetary Property
Ref. Value Weight Exponents
for TSI
for TSI
Mean Radius
Bulk Density
Escape Velocity
Revolution
Surface gravity
1EU
1EU
1EU
1EU
1EU
Surface Temperature
288K
0.57
1.07
0.70
0.70
0.13
0.21
Here, EU = Earth Units, where Earths radius is 6371 km, density is 5.51 g/cm3, escape velocity is 11.19 km/s,
Revolution is 365.25 days, and surface gravity is 9.8 m/s2.
The TSI is classified into Interior TSI and Surface TSI. The interior TSI is the geometrical mean of TSI radius and
density.
TSII = (TSIR × TSIρ)1/2 ,
(7)
4
J. M. Kashyap
Similarly, the surface TSI is defined as the geometrical mean TSI of surface temperature, escape velocity, revolution
of the planet and surface gravity.
The global TSI takes the form:
TSIS = (TSIT × TSIVe × TSIRe × TSIg)1/4
TSI = (TSII × TSIS)1/2 ,
(8)
(9)
Table 4. A sample of calculated TSI
Names
Radius Density Temp E. Vel Rev
g
TSIS
TSII TSI
(EU)
(EU)
Earth
Mars
Kepler-438 b
GJ 667C c
Kepler-296 e
1.00
0.53
1.12
1.54
1.48
1.00
0.73
0.90
1.05
1.03
(K)
288
240
312
280
303
(EU)
1.00
0.45
1.06
1.57
1.50
1.00
1.88
0.09
0.07
0.08
1.00
0.37
1.01
1.61
1.52
1.00
0.84
0.73
0.67
0.69
1.00
0.81
0.96
0.92
0.93
1.00
0.82
0.84
0.78
0.80
Figure 2. Plot of interior NTSI versus surface NTSI. Blue dots are the giant planets, red dots are the rocky planets, and cyan
circles are the Solar System objects (Table 4). However, the optimistic limit for NTSI is marked by a solid line which is ∼ 0.90.
Similar to NESI, the peaks in the plot gives the distribution phenomenon of similarity indices. TSI scatter plot
denotes the optimistic range of each and every gas giants and rocky exoplanets, with respect to Tardigrade survival
ability.
4. DISCUSSION AND CONCLUSION
The Search for Earth-twin is becoming more vibrant area in planetary science research. We know that, ESI was
defined for only four physical parameters. But it is necessary to understand the need for compiling more physical
parameters to find Earth-like planet. Hence, in this paper the NESI has been introduced to get more accurate
Quantitative indexing and Tardigrade analysis of exoplanets
5
results.The results obtained above clarifies it on comparison with ESI with 4 parameters (Schulze-Makuch et al. 2011;
Kashyap et al. 2016). In 2008 (Jnsson et al. 2008), showed that Tardigrade could survive in space for 10 days. Thus
the Tardigrade Similarity Index (TSI) is introduced and analyzed for all 3370 exoplanets, where this extremophile life
could survive. The future work in this area for the quest in search of life can be done by upgrading NESI with more
physical parameters(such as: tilt of the planet, albedo factor, magnetic field, ... etc). Finally, in order to sustain life
atmospheric study in detail plays a major role, therefore analyzing the astrochemistry of the exoplanets is very crucial.
ACKNOWLEDGMENTS
This research has made use of the Extrasolar Planets Encyclopaedia at http://www.exoplanet.eu, Exoplanets
Data Explorer at http://exoplanets.org, the Habitable Zone Gallery at http://www.hzgallery.org/, the NASA
Exoplanet Archive, which is operated by the California Institute of Technology, under contract with the National
Aeronautics and Space Administration under the Exoplanet Exploration Program at
http://exoplanetarchive.ipac.caltech.edu/ and NASA Exoplanet Archive at
http://exoplanetarchive.ipac.caltech.edu and NASA Astrophysics Data System Abstract Service.
REFERENCES
Becquerel, P. (1950). La suspension de la vie au dessous de 1/20
K absolu par demagnetization adiabatique de l'alun de fer
dans le vide les plus elve. C. R. Hebd. Sances Acad. Sci.
Paris, 231, 261.
Jnsson K., Ingemar; Rabbow, Elke; Schill, Ralph O.;
Harms-Ringdahl, Mats and Rettberg, Petra (2008).
Tardigrades survive exposure to space in low Earth orbit.
Current Biology, 18, 729.
Kashyap J. M. (2017). NESI and NTSI, Mendeley Data, v.1,
Brent Y. Creer, Captain Harald A. Smedal, USN (MC), and
DOI: https://data.mendeley.com/datasets/c37bvvxp3zK.
Rodney C.Vtlfngrove (2012). Centrifuge Study of Pilot
Tolerance to Acceleration and the Effects of Acceleration on
Pilot Performance. NASA Technical note D-337.
Copley J. (1999).Indestructible, New Scientist, Retrieved 2010.
Horikawa, Daiki D. (2012). Alexander V. Altenbach; Joan M.
Bernhard; Joseph Seckbach, eds. Anoxia Evidence for
Eukaryote Survival and Paleontological Strategies. (21st ed.).
Springer Netherlands, 205. ISBN 978-94-007-1895-1. Retrieved
2012-01-21.
Kashyap J. M., Safonova M., Gudennavar S. B. (2016). New
indexing and surface temperature analysis of exoplanets,
arXiv:1608.06702.
Schulze-Makuch, D., Mndez, A., Fairn, A. D., et al. (2011a). A
Two-Tiered Approach to Assessing the Habitability of
Exoplanets. Astrobiology, 11, 1041.
Tsujimoto, Megumu; Imura, Satoshi; Kanda, Hiroshi (2015).
Recovery and reproduction of an Antarctic tardigrade
retrieved from a moss sample frozen for over 30 years.
Cryobiology, 72, 78.
|
1711.06372 | 1 | 1711 | 2017-11-17T01:52:05 | Observations of a new stabilizing effect for polar water ice on Mars | [
"astro-ph.EP"
] | Using the Compact Reconnaissance Imaging Spectrometer for Mars (CRISM), we map the temporal variability of water ice absorption bands over the near-polar ice mound in Louth crater, Mars. The absorption band depth of water ice at 1.5 microns can be used as a proxy for ice grain size and so sudden reductions can time any switches from ablation to condensation. A short period of deposition on the outer edge of the ice mound during late spring coincides with the disappearance of seasonal water frost from the surrounding regolith suggesting that this deposition is locally sourced. The outer unit at Louth ice mound differs from its central regions by being rough, finely layered, and lacking wind-blown sastrugi. This suggests we are observing a new stabilizing effect wherein the outer unit is being seasonally replenished with water ice from the surrounding regolith during spring. We observe the transport distance for water migration at Louth crater to be ~4km, and we use this new finding to address why no water ice mounds are observed in craters <9km in diameter. | astro-ph.EP | astro-ph | Observations of a new stabilising effect for polar water
ice on Mars
Adrian J. Brown1*1, Jonathan Bapst2, Shane Byrne2
1 Plancius Research, Severna Park, MD 21146, USA
2Lunar and Planetary Laboratory, University of Arizona, Tucson, AZ, 85721, USA
Using the Compact Reconnaissance Imaging Spectrometer for Mars
(CRISM), we map the temporal variability of water ice absorption bands
over the near-polar ice mound in Louth crater, Mars. The absorption band
depth of water ice at 1.5 microns can be used as a proxy for ice grain size
and so sudden reductions can time any switches from ablation to
condensation. A short period of deposition on the outer edge of the ice
mound during late spring coincides with the disappearance of seasonal
water frost from the surrounding regolith suggesting that this deposition is
locally sourced. The outer unit at Louth ice mound differs from its central
regions by being rough, finely layered, and lacking wind-blown sastrugi.
This suggests we are observing a new stabilizing effect wherein the outer
unit is being seasonally replenished with water ice from the surrounding
regolith during spring. We observe the transport distance for water
migration at Louth crater to be ~4km, and we use this new finding to
address why no water ice mounds are observed in craters <9km in
diameter.
Key Point 1. First observation of Martian process of exchange of perennial water
with the surrounding regolith
Key Point 2. We infer a advection travel distance of ~4km for water ice particles
Key Point 3. Our findings could explain why no craters <9km have water ice
mounds
Corresponding author:
Adrian Brown
Plancius Research
ph. 408 832 6290
email: Adrian.J.Brown@nasa. gov
KEYWORDS
Mars, Polar caps; Mars, climate; Ices; Radiative transfer; Ices, IR spectroscopy
1
~ 1 ~
1. Introduction
The North Polar Residual Cap (NPRC) is a surface water ice deposit that
persists throughout the martian summer (Kieffer 1976) and which unconformably
drapes the much-thicker North Polar Layered Deposits (Byrne 2009; Herkenhoff
et al. 2007; Tanaka et al. 2008). The NPRC is extensive (~106 km2) and plays a
central role as the dominant source and sink of seasonal water in the martian
atmosphere (Smith et al. 2016). Analysis of its albedo (Kieffer 1990) and spectra
(Langevin et al. 2005; Brown et al. 2016) show the NPRC composition is dust-
free large-grained water ice.
Multiple impact craters peripheral to the NPRC contain surface ice with similar
appearance, albedo and spectra (Conway et al. 2012; Brown et al. 2008).
Korolev crater contains the largest of these deposits and has thermal properties
(Armstrong et al. 2005) and underlying layering (Brothers and Holt 2016) that is
similar to the NPRC. Louth crater contains the most equatorial of these deposits
and has been identified as similar in surface texture and composition to the
NPRC (Brown et al. 2008).
Water ice is seasonally added and removed from these deposits in a
repeatable way. Each winter, atmospheric water in the north polar hood is
incorporated into the seasonal frost cap. As this seasonal cap retreats water is
liberated from its edge; however, most of this water recondenses on the
remaining seasonal ice immediately to the north as originally described by
Houben et al. (1997). This Houben-effect results in an annulus of fine-grained
water frost tracking the edge of the seasonal cap as it retreats poleward
(Wagstaff et al. 2008), which sweeps over craters like Louth and Korolev in the
late Spring (Byrne et al. 2008; Brown et al. 2012). During summer the seasonal
frosts disappear and large-grained (old) water ice appears as the NPRC is
exposed. Brown et al. (2016) used CRISM to measure 1.5 μm adsorption band
depth and so identify when fine-grained frost began accumulating in late summer
at Gemini Lingula (Ls~110-130). This reaccumulation causes late summer
brightening of the NPRC observed by Mariner and Viking IRTM (James and
Martin, 1985; Kieffer, 1990; Paige et al., 1994; Bass et al., 2000), TES
(Armstrong et al., 2007; Xie et al., 2008) and MEX-OMEGA (Appere et al., 2011).
The net annual mass balance of the NPRC and its outliers is uncertain. The
exposure of larger-grained (old) ice in the summer indicates that net ablation is
taking place; however, the lack of surface dust indicates that ablation has been
minimal so it's possible that the NPRC has only recently entered this state
(Laskar et al. 2002; Greve et al. 2005). At 70.2N, Louth crater contains the
lowest latitude surface ice on Mars and so it may be even more sensitive to any
recent climatic changes that may have occurred.
Here we study the Louth crater ice mound using the same techniques as
Brown et al. (2016) applied to the NPRC. This convex-up dome-shaped mound is
10 km wide, 250 m thick and displays layering (Conway et al., 2012). Several
formation mechanisms have been proposed for this mound and others like it,
~ 2 ~
including 1) remnants of a more extensive northern ice cap (Garvin et al., 2000;
Tanaka et al., 2008), 2) upwelling from an underground aquifer (Russell and
Head, 2002), 3) melting and ponding of near surface ice by impact induced
hydrothermal activity (Rathbun and Squyres, 2002) and 4) atmospheric direct
deposition as individual outliers (Brown et al., 2008; Conway et al., 2012). We
examine the seasonal timing of switches between accumulation and ablation,
compare these observations to recent simulations of Bapst et al. (2017) and
discuss the implications of our results for the formation of the ice mound and its
long term fate.
2. Methods
CRISM is a visible to near-infrared imaging spectrometer with 545 channels
from 0.365-3.94 μm (Murchie et al., 2007). In this work, we utilize Full Resolution
Targeted (FRT: 545 channels, 18 m/pix), Half resolution long (HRL: 545
channels, 36 m/pix) and Multispectral Mapping (MSP: 72 channels, 180m/pix)
observations. CRISM observations are ~10km wide, but vary in length by
observation type.
The depth of the 1.5 μm water ice absorption band is a non-linear proxy for ice
grain size (Langevin et al. 2005; Brown et al., 2012, 2014, 2016) and is defined
by an H2O index:
H 2Oindex=1 −
R (1.5)
R (1.394)0.7 R (1.75)0.3
(1)
where R(x) is the reflectance at wavelength x. This index saturates at values of
~0.7 or grain sizes of ~100 μm. Increases in the H2O index (grain size) may
occur by removal or thermal metamorphism of fine-grained ice, whereas
decreases are commonly associated with condensation or precipitation of fresh
finer-grained water ice.
We calculated the H2O index for each CRISM pixel and used the "MR PRISM"
software suite (Brown et al., 2004; Brown et al., 2005; Brown and Storrie-
Lombardi, 2006) to map-project these images into a polar stereographic space
with a common pixel size.
3. Results
Figure 1 shows three CTX images of Louth Crater from late northern summer
in Mars Year (MY) 30. The central and right images show late summer images
with low albedo markings on the ice mound (red arrows). Similar dark markings
that have a streaky appearance have also appeared on the pole facing crater rim
(blue arrows).
~ 3 ~
3.1 H2O index maps
The H2O index for four periods during mid northern summer in MY29, 30 and
31 are shown in Figure 2a-c, to highlight changes in the water ice index over the
Louth ice mound during this period. Figures 2a-c show water ice index images
with a stretch that is designed to show small changes in the index on the lower
end of the scale. This has the effect of bringing out subtle changes in water ice
index in the regolith around the Louth ice mound. This stretch also allows us to
observe that during the period, following the retreat of the seasonal CO2 ice cap,
there is a large halo of water ice on the crater floor, surrounding the ice mound,
which is either deposited during the previous fall before the CO2 seasonal cap
was deposited, or was left over as a residuum from the seasonal cap retreat due
to the higher stability of water ice compared to CO2 ice (Titus, 2005; Titus et al.
2008; Wagstaff et al., 2008; Brown et al., 2012). A key question we can now
address is what is the fate of this regolith-covering water ice as temperatures
warm?
Figure 3 displays the MY30 dataset with a stretch that is intended to bring out
changes on the high side of the index (corresponding to large grain sizes). This
shows the increase in H2O index, particularly in the Ls=108 image, in the center
of the mound. Changes around the edge of the mound are much more muted
(they appear to be consistently green), however the white arrows indicate the H2O
index on the southern edge of the mound increases significantly from Ls=59 to
108.
The five H2O index maps for MY29 covering the Ls=62-149 period are shown
in Figure 2a. These show the gradual increase of the H2O index over the ice
mound over this time. The last image at Ls=149 is almost uniformly white
(indicated by a pink arrow). The images also show a decrease in the H2O index in
the regolith around the ice mound. In particular, it should be noted that at Ls=62-
67, around the southern rim of the ice mound, a halo of water ice is present (this
is colored light blue and indicated by a white arrow in the Ls=62 image). This
disappears markedly in the Ls=92-149 period. The last two images show strong
H2O index values in an annulus which is the rim of the crater to the south of the
ice mound.
The MY30 H2O ice index maps are shown in Figure 2b over the period from
Ls=59, when seasonal CO2 is finally disappearing, to Ls=108. It should be noted
that the first image at Ls=59 shows a relatively high H2O index on the regolith
around the edge of the cap (this is colored green and is indicated by a white
arrow). This large ice index is largely gone by the Ls=83 image, similar to the
previous Mars Year. This is colored red and indicated by a green arrow on the
Ls=83 image.
The MY31 H2O index images covering Ls=75-165 are shown in Figure 2c.
This image sequence again shows an increase in H2O index on the interior of the
~ 4 ~
mound. The possible exception is for the late summer Ls=92 image, which
appears to have a slightly lower H2O index on the north edge of the water ice
mound (indicated by a pink arrow in the image). We do not have a satisfactory
explanation for this observation. As for the previous two years, the first image
shows the highest H2O index on the regolith surrounding the ice mound. The
Ls=75 image shows a spatial variation in the water ice index, being lower around
the edge of the mound (this is indicated by a white arrow).
Figure 3 displays the MY30 dataset with a stretch that is intended to bring out
changes on the high side of the index (corresponding to large grain sizes). This
shows the increase in H2O index, particularly in the Ls=108 image, in the center
of the mound. Changes around the edge of the mound are much more muted
(they appear to be consistently green), however the white arrows indicate the H2O
index on the southern edge of the mound increases significantly from Ls=59 to
108.
3.3 Difference images
We constructed difference images for four of the images of Figure 2a, and
these are shown in Figure 4a. Only the overlapping pixels in each image are
processed and presented. Co-registration of the images was carried out as
described in Section 2.1.1. The left panel of Figure 4a shows a slight decrease in
the H2O ice index between Ls=65 and 67, colored in orange and indicated by a
pink arrow. The right panel of Figure 4a shows a dark rim around the edge of the
crater, indicating a large decrease in the H2O index around the rim. The affected
area is colored in black and indicated by a white arrow. The other feature of note
is a speckle of blue pixels to the south of the crater in the regolith. These areas
also experienced a decrease in the H2O index between Ls=67 and 92.
Figure 4b shows the three difference images between the three images of
Figure 2b calculated in the same manner as Figure 4a. Most strikingly, Figure 4b
shows that in Ls=59-83, there is a decrease in H2O ice index in the regolith
surrounding the ice mound (these areas are colored green and blue and
indicated by a pink arrow), particularly on the south side of the mound. In the
period following this (from Ls=83-91), there is a corresponding increase in the H2O
index on the south side of the ice mound (this region is colored white and the
location indicated by grey arrows). The spatial occurrence of the increased H2O
index on the south side of the mound and decrease in the water ice in the
adjacent regolith in the following period are significant and the possible causes of
this observation will be addressed in the Discussion section below.
Comparing Figure 4a and 4b, we can make a direct comparison between the
second panel of 4a (Ls=67-92) and the first panel of 4b (Ls=59-83). The time
periods are slightly different, which makes a significant difference, and in addition
to this interannual differences will play a confounding role here. However, there
are some similarities which lead us to believe the behavior we observe is a
~ 5 ~
regular occurrence. For example, the interior of the water ice mound is white in
both images, indicating increasing of H2O index in this region. The H2O ice index
of the surrounding regolith is blue in some locations, indicating a decrease in
those regions. The rim of the mound in Figure 4a (MY 29 Ls=67-92) is black,
indicating a strong decrease in the H2O index along the rim at this time, in
addition, the rim of the mound in Figure 4b (MY 30 Ls=59-83) changes from white
to red, indicating a small decrease at that time. Therefore we believe similar
trends exist in this key period in MY 29 (Figure 4a) and MY 30 (Figure 4b).
3.4 Average and maximal H2O ice index
Figure 5 shows a gridded spatial and temporal sequence that captures each
the H2O index maps of each CRISM image that obtained useable data over the
Louth ice mound (see also Table 1). The images are arranged as function of time
in solar longitude (from Ls=59 to 149, left to right) for Mars Year 29, 30 and 31,
from top to bottom. Note that the temporal (left-right) scale is not linear, we have
simply ordered the images from left to right and attempted to line up images of
the same time in a Mars Year.
Using this gridded time sequence, we can more easily perceive three points
that are critical to establishing the intra and inter-annual behavior of ice on the
Louth ice mound. First, the amount of water ice on the surrounding regolith
decreases from Ls=59 to Ls=165, having reached a stable minimum amount by
about Ls=92.
The second point to note is that the water ice index on the water ice mound
interior increases relatively steadily over the first two Mars Years (29, 30),
however in one Ls=80 observation in Mars Year 31, the water ice index increases
and then decreases in Ls=92. This phenomenon is not as strongly apparent in the
other Mars Years. We address this point further in the Discussion section below.
The third major point to note is the decease in the water ice index around the
edge of the mound, particularly on the north side around Ls=92 in each Mars Year
(this is colored red and indicated with green arrows) but also on the south side of
the ice mound (indicated by pink arrows). These decreases are likely caused by
deposition of fine grained water ice around the edge of the ice mound – this
observation is critical to our model, and is addressed further in the Discussion
section.
The late summer CRISM image MSP D25A at Ls=149 shows the highest H2O
index values (this image is also shown in Figure 3a). There is an indication of a
turn over in the MSP 314FD observation of MY30 at Ls=165 (see also Figure 2c),
although this could also be due to the onset of the polar hood obscuring the
crater at this time (Benson et al., 2011; Brown et al., 2016).
~ 6 ~
3.5 HiRISE spring and summer images of Louth
In order to further our understanding of the regolith-mound water ice
exchange process, Figure 6 shows two HiRISE images taken at Ls=75.1 and
Ls=156. The close up images in the bottom of the image show the rough "stucco"
nature of the exterior unit of the mound. This was stucco unit was mapped as
"Unit 4" by Brown et al. (2008). Two key morphological observations are
important to note in these HiRISE images. The first is that the northern rim of the
ice mound contact with the regolith is clean and relatively distinct (blue arrows in
Figure 6). The southern rim of the ice mound is ragged and less distinct (red
arrows). This is a strong indication of southerly winds transporting water ice in the
northerly direction, because the ragged edge indicates the direction from which
ice is removed, and the sharp edge shows the direction in which ice has newly
covered regolith. This is in accord with the asymmetric offset of the ice mound
within Louth.
The second observation to make regarding these HiRISE images is the linear
features in the middle of the central ice mound. These were interpreted to be
"sastrugi" by Brown et al. (2008) and found to be present in their "Unit 1, the
smooth undulating layer in the center of the Mound", and less prominently in their
mapped "Unit 3". These features run almost north-south (see black lines in Figure
6), which is in accord with their interpretation as wind-blown sastrugi, which on
Earth run parallel to the wind direction (Gray and Male, 1981). This is roughly to
the north-north east, in accord with the sharp (blue arrows) and ragged (red
arrows) edges discussed above.
4. Discussion
4.1 Seasonal evolution of water frosts
It is clear from Figures 4 and 5 that the seasonal behavior of the mound is not
uniform and is quite different from the surrounding regolith. The water ice index
over the regolith decreases markedly from ~0.4 to ~0.1 during Ls=59 to 92 (white
arrows in Figure 2a and b), and then stabilizes at these low values (dark blue in
figures) indicating the disappearance of a fine-grained seasonal water frost. In
contrast, the water ice index on the ice mound generally starts higher (~0.65 at
Ls=80 in MY31) and then undergoes a slight decrease (to ~0.6) at the edges of
the ice mound from Ls=83-91 (green and pink arrows in Figure 5). Thus, the
timing of this decrease in water ice index at the mound edge corresponds to the
removal of seasonal water frost on the surrounding regolith.
We suggest that water sublimated from the regolith recondenses on the
periphery of the (much colder) mound. The spatial pattern of this late-spring
condensation is crescent-shaped (Figure 4a), being wider on the southwest edge
of the mound. Winds from the southwest could advect regolith-derived water
vapor over the mound. As vapor passes over this colder surface it condenses
(lowering the water ice index); however, after traversing ~4km of this icy surface
~ 7 ~
the humidity of the air has fallen to the point where further condensation does not
occur (leaving the mound interior frost-free). This is similar to the 'Houben' effect
that plays out at larger scales at the edge of the retreating seasonal cap (Houben
et al. 1997; Wagstaff et al. 2008).
Similar mechanisms have also been
suggested to create albedo patterns on the North Polar Residual Cap that could
ultimately lead to the formation of troughs (Ng and Zuber, 2006).
In addition to its differing water ice index evolution, the periphery of the mound
exhibits distinct geomorphology. Figure 9 of Brown et al. (2008) mapped four
different units within the water ice using CRISM and HiRISE observations,
including an outer finely-layered 'stucco rough ice' unit and an inner, sastrugi-
bearing 'smooth interior unit'.
The outer unit corresponds to the area that
experiences recondensation of regolith-supplied water vapor and, over time, this
additional seasonal frost exchange may lead to its distinct morphology.
4.2 Interannual comparisons
Brown et al. (2008) identified small ice outliers in HiRISE imagery next to the
edge of the Louth ice mound and concluded that it may be in retreat. A multi-
year comparison of the edge location of the ice mound showed no detectable net
change however (Bapst et al. 2017); although, some interannual variability was
seen.
Figure 5 shows a significant difference in the H2O ice index between MY 29
and 30. There was a global dust storm in winter of MY 28 (Kass et al., 2007)
when Louth crater was covered with seasonal CO2 ice. The dust storm transiently
warmed the Martian climate system, significantly affected the late-summer
appearance of the south polar residual cap in MY28 (Becerra et al., 2015), and
likely led to faster springtime recession of the northern seasonal cap in MY 29
(Piqueux et al., 2015). The unusually low H2O index during spring of MY29 may
be linked to the climatic disruption caused by this dust storm.
4.3 Comparison to 1D depositional model results
Bapst et al. (2017) discuss the details of a 1-D depositional model to constrain
the mass balance of the ice mounds at Louth and Korolev. This model
approximates the Louth Crater ice mound as an infinite ice sheet with a constant
water-ice albedo that can be varied for different runs. The results are therefore
most applicable for the interior of the mound and do not capture the edge water
ice exchange effects discussed (e.g. in Figure 5). Figure 7 shows the results of
this model for Louth, for a range of albedo (0.35-0.43), and for the nominal
windspeed (3 m s-1). For the albedo range shown, the model predicts a short
period of accumulation of water ice, followed by a longer period of sublimation
during northern summer. This range of models all result in net-annual ablation,
with ranging from 3.6 to 0.1 mm of water ice lost annually, assuming an ice
density of 920 kg m-3.
~ 8 ~
Of special relevance to this paper is the short period of accumulation of water
ice after the CO2 seasonal cap has disappeared, occurring from Ls 60-90, and
interpreted as cold-trapping by the water ice mound. Prior to Ls ~80,
temperatures are too low to result in net sublimation of water ice. However,
during northern summer, the ice mound undergoes net sublimation. The
observed transition from accumulation to ablation, on the outer part of the
mound, occurs between LS 92-100 in MY29, LS 83-91 in MY30 and LS>92 in
MY31 (Figure 5). From these constraints, we estimate the timing of this transition
between Ls 83-100 and that the mound continues to ablate for the remainder of
the summer. The timing of this transition is matched by models with water ice
albedos of 0.37-0.41 (Figure 7). The modeled albedo is not necessarily
representative of the actual ice deposit, but shows the sensitivity of this
parameter and how it can be tested against CRISM observations of water-ice
index.
Using this albedo range of 0.37-0.41, the net-annual change predicted from
the model of Bapst et al. (2017) are 2.4 to 0.7 mm of water ice loss per Mars
year. Thus, the model (constrained by early summer observations) and the late-
summer CRISM and CTX data paint a consistent picture of the ongoing net-
annual loss of ice from the Louth crater mound.
4.4 Implications for NPRC and origin of outlying water ice mounds
The observations of water ice transport to the Louth ice mound from the
surrounding regolith in the Ls=59-91 period (Figure 2b) may have implications for
the NPRC and water ice outliers around the NPRC, in addition to other crater
hosted ice mounds such as those in Korolev and Dokka craters. Figure 8 shows
an interpretive diagram of the transfer process through spring suggested by our
observations. This diagram shows the following key events in the model: 1) Ls=0-
55, Louth Crater ice mound is covered by seasonal CO2 ice and is inactive. 2)
Ls=55-70, Water ice is exposed on the regolith surrounding the ice mound. 3)
Ls=80-90, Water ice transfers from the regolith to the mound in the Louth crater
situation. A similar process would also likely take place at the edge of large ice
mounds craters like Korolev, other water ice outliers and also the edge of the
NPRC.
We find it intriguing that there is an apparent advection distance for water
transport from the regolith to the Louth ice mound of ~4km. If this transport
distance observed at Louth is applicable to other crater hosted ice mounds, we
can use this information to speculate on a previously unexplained polar
phenomenon – that no craters less than 9km in diameter host water ice deposits
(Conway et al., 2012). Since the travel distance of water by advection in the
Louth crater environment is approximately 4km, which is remarkably similar to
the observed lower limit of 9km for craters with ice mounds. This may indicate
that a protected zone like a crater may have to have a radius that exceeds at
least twice this distance (8km) in order for this process to be form and preserve
~ 9 ~
this type of perennial ice deposit.
Our observation of an apparent advection distance of ~4km could suggest that
for ice mounds to form (and perhaps remain stable), the ice mounds require a
certain size and amount of "free space" in which they can interact with the
surrounding regolith and cold trap advected water them that can be transported
over km size distances. In the case of Louth, the ice mound is probably protected
by the exterior smooth unit that, as we have reported, is fed and likely dominated
by regolith bound water. This might be the key to its relatively large size and
stability, in spite of its distance from the pole.
Considering the scenario of a smaller crater, one might envision that the water
ice that is deposited in the bottom of it during winter might travel kilometers by
this spring advection transport mechanism to the crater rims and get trapped
there, rather than forming a central mound. This may help explain why no ice
mounds have formed in craters <9km in diameter.
This potential explanation is speculative at this stage and requires numerical
studies and further polar observations. One powerful observational method
would be mapping by an active multispectral lidar, with polarization capability, to
discriminate cloud composition, aerosols and dust, as recently discussed in
(Brown et al., 2015) and highlighted in the recent study by Heavens (2017).
5. Conclusions
We have used CRISM data to map the spatial and temporal behavior of water
ice within Louth crater. Data from three Mars years show a seasonal increase of
the H2O index in the interior of the mound and a decrease on the surrounding
regolith.
The significant findings of our study are:
1. We have established for the first time that the Louth crater ice mound ablates
throughout summer, at least until Ls=150 (Figure 5).
2. We established that the outer edge of the Louth water ice mound exchanges
water ice with the surrounding regolith. As the region warms in spring, water
ice sublimates from the regolith surface and is then cold trapped as fine-
grained ice onto the edge of the water ice mound. This fine-grained water ice
exchange may contribute to the fine layering observed by Conway et al
(2012) and the texture of the external unit of the ice mound identified by
Brown et al. (2008).
3. The spatial pattern of frost formation on the mound suggests that
southwesterly winds are responsible for carrying water vapor across the
mound leading to a ~4km wide zone of temporary frost formation. This finding
may play a role in explaining the observation that no craters <9km in diameter
posses water ice mounds.
4. The surrounding regolith is usually neglected in 1D models when calculating
~ 10 ~
stability of exposed surface ice on Mars (Bapst et al. 2017; Dundas and Byrne
2010). This work shows that surrounding regolith can act as an enhanced
source of water vapor in some seasons and should be included in future
assessments of ice stability.
6. References
Appere, T., Schmitt, B., Langevin, Y., DoutÈ, S., Pommerol, A., Forget, F., Spiga, A.,
Gondet, B., Bibring, J.P., 2011. Winter and spring evolution of northern seasonal
deposits on Mars from OMEGA on Mars Express. J. Geophys. Res. 116, E05001.
Armstrong, J.C., Nielson, S.K., Titus, T.N., 2007. Survey of TES high albedo events in
Mars' northern polar craters. Geophys. Res. Lett. 34.
Armstrong, J.C., Titus, T.N., Kieffer, H.H., 2005. Evidence for subsurface water ice in
Korolev crater, Mars. Icarus 174, 360–372.
Bapst, J., Bandfield, J.L., Wood, S.E., 2015. Hemispheric Asymmetry in Martian
Seasonal Surface Water Ice from MGS TES. Icarus 260, 396–408.
doi:10.1016/j.icarus.2015.07.025
Bapst, J., Byrne, S., 2016. Louth Crater Water Ice as a Martian Climate Proxy. Presented
at the LPSC XXXXVII, LPI, Houston, TX, p. Abstract #3027.
Bapst, J., Byrne, S., Brown, A.J., 2017. On the Icy Edge at Louth and Korolev Craters.
Icarus, 10.1016/j.icarus.2017.10.004
Bass, D.S., Herkenhoff, K.E., Paige, D.A., 2000. Variability of Mars' North Polar Water
Ice Cap: I. Analysis of Mariner 9 and Viking Orbiter Imaging Data. Icarus 144,
382–396.
Becerra, P., Byrne, S., Brown, A.J., 2015. Transient Bright "Halos" on the South Polar
Residual Cap of Mars: Implications for Mass-Balance. Icarus 211–225.
doi:10.1016/j.icarus.2014.04.050
Benson, J.L., Kass, D.M., Kleinbhl, A., 2011. Mars' north polar hood as observed by the
Mars Climate Sounder. J. Geophys. Res. 116, E03008.
Brothers, T.C., Holt, J.W., 2016. Three-dimensional structure and origin of a 1.8 km thick
ice dome within Korolev Crater, Mars. Geophys. Res. Lett. 43, 1443–1449.
Brown, A. J., T.J. Cudahy, and M.R. Walter. "Short Wave Infrared Reflectance
Investigation of Sites of Palaeobiological Interest: Applications for Mars
Exploration." Astrobiology 4, no. 3 (2004): 359–76.
https://doi.org/10.1089/ast.2004.4.359.
Brown, Adrian J., M.R. Walter, and T.J. Cudahy. "Hyperspectral Imaging Spectroscopy of
a Mars Analog Environment at the North Pole Dome, Pilbara Craton, Western
Australia." Australian Journal of Earth Sciences 52, no. 3 (2005): 353–64.
https://doi.org/10.1080/08120090500134530.
Brown, A.J., Byrne, S., Tornabene, L.L., Roush, T., 2008. Louth Crater: Evolution of a
layered water ice mound. Icarus 196, 433–445.
Brown, A.J., Calvin, W.M., Becerra, P., Byrne, S., 2016. Martian north polar cap summer
water cycle. Icarus 277, 401–415. doi:10.1016/j.icarus.2016.05.007
Brown, A.J., Calvin, W.M., Murchie, S.L., 2012. Compact Reconnaissance Imaging
Spectrometer for Mars (CRISM) north polar springtime recession mapping: First
~ 11 ~
3 Mars years of observations. J. Geophys. Res. 117, E00J20.
Brown, A.J., Michaels, T.I., Byrne, S., Sun, W., Titus, T.N., Colaprete, A., Wolff, M.J.,
Videen, G., 2015. The science case for a modern, multi-wavelength, polarization-
sensitive LIDAR in orbit around Mars. J. Quant. Spectrosc. Radiat. Transf. 131–
143.
Brown, A.J., Piqueux, S., Titus, T.N., 2014. Interannual observations and quantification
of summertime H2O ice deposition on the Martian CO2 ice south polar cap. Earth
Planet. Sci. Lett. 406, 102–109. doi:10.1016/j.epsl.2014.08.039
Brown, A.J., Storrie-Lombardi, M.C., 2006. MR PRISM - A spectral analysis tool for the
CRISM. Presented at the Proceedings of SPIE Optics and Photonics, SPIE, San
Diego, CA, p. Abstract 28. doi:10.1117/12.677107
Conway, S.J., Hovius, N., Barnie, T., Besserer, J., LeMouélic, S., Orosei, R., Read,
N.A., 2012. Climate-driven deposition of water ice and the formation of mounds
in craters in Mars' North Polar Region. Icarus 220, 174–193.
doi:http://dx.doi.org/10.1016/j.icarus.2012.04.021
Garvin, J.B., Sakimoto, S.E.H., Frawley, J.J., Schnetzler, C., 2000. North polar region
craterforms on Mars: Geometric characteristics from the Mars Orbiter Laser
Altimeter. Icarus 144, 329–352.
Gray, D.M., Male, D.H., 1981. Handbook of Snow: Principles, Processes, Management
and Use. Pergamon, Oxford.
Greve, Ralf, and Rupali A. Mahajan. "Influence of Ice Rheology and Dust Content on the
Dynamics of the North-Polar Cap of Mars." Icarus 174, no. 2 (2005): 475–85.
Heavens, N.G., 2017. The reflectivity of Mars at 1064 nm: Derivation from Mars Orbiter
Laser Altimeter data and application to climatology and meteorology. Icarus 289,
1–21. doi:10.1016/j.icarus.2017.01.032
Herkenhoff, K.E., S. Byrne, P.S. Russell, K.E. Fishbaugh, and A.S. McEwen. "Meter-
Scale Morphology of the North Polar Region of Mars." Science 317, no. 5845
(2007): 1711–15. DOI: 10.1126/science.1143544.
Houben, H., Haberle, R.M., Young, R.E., Zent, A.P., 1997. Modeling the Martian
seasonal water cycle. J. Geophys. Res.-Planets 102, 9069–9083.
Jakosky, B.M., 1985. The seasonal cycle of water on Mars. Space Sci. Rev. 41, 131–200.
James, P.B., Martin, L.J., 1985. Interannual water loss by Mars' north polar cap., in:
Bulletin of the American Astronomical Society. p. 735.
Kass, D.M., McCleese, D., Schofield, T., Kleinboehl, A., Zurek, R., Bowles, N., 2007.
MCS views of the 2007 global dust storm, in: Bulletin of the American
Astronomical Society. p. 441.
Kieffer, H., 1990. H2O grain size and the amount of dust in Mars' residual north polar
cap. J. Geophys. Res. 95, 1481–1493.
Kieffer, H.H., Chase, S.C., Martin, T.Z., Miner, E.D., Palluconi, F.D., 1976. Martian
North Pole Summer Temperatures: Dirty Water Ice. Science 194, 1341–1344.
Langevin, Y., Bibring, J.-P., Montmessin, F., Forget, F., Vincendon, M., Douté, S., Poulet,
F., Gondet, B., 2007. Observations of the south seasonal cap of Mars during
recession in 2004–2006 by the OMEGA visible/near-infrared imaging
spectrometer on board Mars Express. J. Geophys. Res. 112,
10.1029/2006JE002841.
Langevin, Y., Poulet, F., Bibring, J.-P., Schmitt, B., Doute, S., Gondet, B., 2005. Summer
~ 12 ~
Evolution of the North Polar Cap of Mars as Observed by OMEGA/Mars
Express. Science 307, 1581–1584.
Laskar, J., Levrard, B., Mustard, J.F., 2002. Orbital forcing of the martian polar layered
deposits. Nature 419, 375–377.
Murchie, S., Arvidson, R., Bedini, P., Beisser, K., Bibring, J.-P., Bishop, J., Boldt, J.,
Cavender, P., Choo, T., Clancy, R.T., Darlington, E.H., Des Marais, D., Espiritu,
R., Fort, D., Green, R., Guinness, E., Hayes, J., Hash, C., Heffernan, K.,
Hemmler, J., Heyler, G., Humm, D., Hutcheson, J., Izenberg, N., Lee, R., Lees, J.,
Lohr, D., Malaret, E., T., M., McGovern, J.A., McGuire, P., Morris, R., Mustard,
J., Pelkey, S., Rhodes, E., Robinson, M., Roush, T., Schaefer, E., Seagrave, G.,
Seelos, F., Silverglate, P., Slavney, S., Smith, M., Shyong, W.-J., Strohbehn, K.,
Taylor, H., Thompson, P., Tossman, B., Wirzburger, M., Wolff, M., 2007.
Compact Reconnaissance Imaging Spectrometer for Mars (CRISM) on Mars
Reconnaissance Orbiter (MRO). J. Geophys. Res. 112, E05S03,
doi:10.1029/2006JE002682.
Ng, Felix S. L., and Maria T. Zuber. "Patterning Instability on the Mars Polar Ice Caps."
Journal of Geophysical Research: Planets (1991–2012) 111, no. E2 (February 1,
2006). doi:10.1029/2005JE002533.
Paige, D.A., Bachman, J.E., Keegan, K.D., 1994. Thermal and albedo mapping of the
polar regions of Mars using Viking thermal mapper observations: 1. North polar
region. J. Geophys. Res. 99, 25959–25991.
Piqueux, S., Kleinböhl, A., Hayne, P.O., Kass, D.M., Schofield, J.T., McCleese, D.J.,
2015. Variability of the martian seasonal CO 2 cap extent over eight Mars Years.
Icarus 251, 164–180.
Putzig, N.E., Phillips, R.J., Campbell, B.A., Holt, J.W., Plaut, J.J., Carter, L.M., Egan,
A.F., Bernardini, F., Safaeinili, A., Seu, R., 2009. Subsurface structure of Planum
Boreum from Mars Reconnaissance Orbiter Shallow Radar soundings. Icarus 204,
443–457.
Rathbun, J.A., Squyres, S.W., 2002. Hydrothermal systems associated with martian
impact craters. Icarus 157, 362–372.
Richardson, M.I., Wilson, R.J., 2002. A topographically forced asymmetry in the martian
circulation and climate. Nature 416, 298–301.
Russell, P.S., Head, J.W., 2002. The martian hydrosphere/cryosphere system:
Implications of the absence of hydrologic activity at Lyot crater. Geophys. Res.
Lett. 29, art. no.-1827.
Smith, I.B., Putzig, N.E., Holt, J.W., Phillips, R.J., 2016. An ice age recorded in the polar
deposits of Mars. Science 352, 1075–1078. doi:10.1126/science.aad6968
Tanaka, K.L., Rodriguez, J.A.P., Skinner Jr., J.A., Bourke, M.C., Fortezzo, C.M.,
Herkenhoff, K.E., Kolb, E.J., Okubo, C.H., 2008. North polar region of Mars:
Advances in stratigraphy, structure, and erosional modification. Icarus, Mars
Polar Science IV 196, 318–358. doi:10.1016/j.icarus.2008.01.021
Titus, T.N., 2005. Thermal infrared and visual observations of a water ice lag in the Mars
southern summer. Geophys. Res. Lett. 32, doi:10.1029/2005GL024211.
Titus, T.N., Calvin, W.M., Kieffer, H.H., Langevin, Y., Prettyman, T.H., 2008. Martian
polar processes, in: Bell, J.F. (Ed.), The Martian Surface: Composition,
Mineralogy, and Physical Properties. Cambridge University Press, pp. 578–598.
~ 13 ~
Wagstaff, K.L., Titus, T.N., Ivanov, A.B., Castano, R., Bandfield, J.L., 2008.
Observations of the North Polar Water Ice Annulus on Mars using THEMIS and
TES. Planet. Space Sci. 56, 256–265.
Xie, H., Guan, H., Zhu, M., Thueson, M., Ackley, S.F., Yue, Z., 2008. A conceptual model
for explanation of Albedo changes in Martian craters. Planet. Space Sci. 56, 887–
894.
7. Acknowledgements
This work was sponsored by NASA MDAP Grant number NNX16AJ48G. We
would like to acknowledge the work of the CRISM science operations team at
JHU APL who acquired this wonderful dataset. We would like to thank two
anonymous reviewers for their helpful comments. We would also like to thank the
creators of The Expanse for key inspirations to this project, and thank in advance
the future members of the MCRN for their dedication to duty and for daring to
dream of a bright and wet future for Mars.
CRISM ID
CRISM
Mode
Earth
DOY
Ls
%
Coverage
Mars Year 28
2F70
3037
3F8E
3B9C
7F89
Mars Year 29
8E90
92AB
9474
9654
99FE
9C9C
A266
A3BE
A84E
A928
AB68
FRT
MSP
FRT
MSP
MSP
MSP
FRT
FRT
FRT
FRT
FRT
FRT
FRT
FRT
HRS
FRT
2006 315 133
2006 320 136.4
2006 340 146.43
2007 003 160.38
2007 273 323
~50
0
25
90
90
10
50
25
50
25
95
90
25
25
75
90
2007 357 7
2008 003 12
2008 008 14.6
2008 013 17
2008 025 22.37
2008 036 27.61
2008 058 37.92
2008 064 40.24
2008 109 60.25
2008 114 62.51
2008 125 67.44
~ 14 ~
B4B1
B965
D25A
Mars Year 30
152F1
166A4
16FDB
173B2
17994
185A3
18879
18CCF
196E6
1B7F0
1C6FD
Mars Year 31
232E0
233FB
237CE
242CD
263D8
Mars Year 32
2ED28
314FD
FRT
MSP
MSP
MSP
HRL
FRT
MSP
FRT
HRL
FRT
HRS
HRS
MSP
HRS
MSP
FRT
HRL
FRT
MSP
MSP
MSP
2008 182 92.20
2008 198 99.6
2008 303 149.46
2010 002 32.46
2010 040 49.55
2010 063 59.54
2010 074 64.46
2010 090 71.44
2010 118 83.34
2010 124 86.05
2010 135 91.03
2010 173 108.05
2010 294 167.91
2010 350 199.43
2012 051 73.20
2012 055 74.94
2012 067 79.75
2012 094 91.62
2012 193 137.94
2014 086 108.20
2014 202 165.24
95
91
99
80
0
95
0
95
90
75
90
75
90
50
0
65
90
65
3
60
40
Table 1 – CRISM observations of Louth crater used in this study. The CRISM
modes (which are discussed more fully in the text) are FRT= Full Resolution
Targeted, HRL = Half Resolution Targeted, HRS = Half Resolution Short, MSP =
Multispectral mapping. DOY = Day of Year.
~ 15 ~
Figure 1 - Louth Crater as captured by MRO MSSS Context (CTX) camera in
Mars Year 30 at Ls=134.1, 167.8 and 183.2, showing the development of an
unexplained dark pattering (red arrows) in late summer to early-fall. North is up.
The CTX instrument image identifiers are G02_019013_2503_XN_70N257W,
G04_019857_2503_XN_70N257W, G05_020213_2503_XN_70N256W. The dark
markings are largely limited to the ice mound, however some have developed on
the ice covered pole-facing rim (blue arrows).
~ 16 ~
Figure 2a - Mars Year 29 CRISM observations of the evolution of the H2O index
for late spring and early summer for Louth crater ice mound and the surrounding
regolith. North is up. At Ls=62, water ice is seen on the regolith surrounding the
water ice mound (white arrow). High H2O indexes correspond to larger grained
water ice, primarily seen in the interior of the ice mound, although by Ls=149, the
water ice index is uniformly high across the water ice mound (pink arrow),
indicating the presence of relatively large grained water ice across the mound by
that late summer period. The blue "bite" out of the north east corner of the ice
mound corresponds to the ice-poor dark sand dunes at this location (see also
Figure 1).
~ 17 ~
Figure 2b – Mars Year 30 H2O ice index maps of Louth Crater for northern spring
and summer. North is up. At Ls=59, there is a large amount of water ice on the
regolith around the water ice mound (white arrow). At Ls=83, the H2O ice index
decreases at the edge of the water ice mound (this region is red and indicated by
a green arrow). At Ls=91 and 108, there is a relatively uniform high water ice
index across the ice mound.
~ 18 ~
Figure 2c – Mars Year 31 H2O ice index maps of Louth Crater for northern spring
and summer. North is up. At Ls=75, there is a relative low on the north edge of
the water ice mound (this area is colored red and indicated by a white arrow). At
Ls=80, the mound is relatively high and uniform across the mound (this is
indicated in white). At Ls=92, the water ice index has decreased again on the
north side of the ice mound (indicated by the pink arrow), for which we do not
have a satisfactory explanation. At Ls=165 there is insufficient coverage of the
mound to make firm conclusions regarding the water ice index spatial distribution
across the mound. The water ice index of the crater rim (south east of the
mound) is seen to be moderate in this image (it is colored red, yellow and green).
~ 19 ~
Figure 3 – H2O index image for Louth Crater for MY 30. This image is the same
as Figure 2b, but stretch is altered to show the relative difference of the H 2O ice
index within the water ice mound. North is up. The H2O index of central part of
the mound and the southern edge (arrows) are seen to brighten significantly from
Ls=59-108.
~ 20 ~
Figure 4a - Difference images for three images of Figure 3a, showing change in
H2O index over summer for water ice mound and surrounds. North is up. Note
the small decrease in water ice in the regolith from Ls=62-67 (this is colored
orange and indicated by a pink arrow) and the large decrease in the H2O ice
index on the edge of the mound from Ls=67-92 (this is colored black and
indicated by a white arrow). The decrease in the H2O ice index is due to
deposition of fine grained water ice, which has the effect of decreasing the depth
of the H2O ice absorption band. See the text for further details.
~ 21 ~
Figure 4b - Difference between four images in Figure 3b, showing change in H2O
index over summer for water ice mound and surrounds. North is up. Note in
Ls=59-83, there is a decrease in H2O ice index in the regolith surrounding the ice
mound (pink arrow), particularly on the south side of the mound. Note also in
Ls=83-91 there is a corresponding increase in the H2O index on the south side of
the ice mound (grey arrows).
~ 22 ~
Figure 5 –H2O ice index images at Louth crater as a function of solar longitude
(from Ls=59-149, left to right) for Mars Year 29, 30 and 31. Note that the temporal
scale is not linear, the images are time ordered from left to right. Points to note:
1.) The amount of water ice on the surrounding regolith decreases from Ls=59 to
Ls=165, and reaches a stable minimum amount by Ls=92. 2.) The water ice index
on the water ice mound increases relatively steadily over the first two Mars Years
(29, 30), however in one Ls=80 observation in Mars Year 31, the water ice index
increases then decreases in Ls=92. This phenomenon is not as strongly apparent
in the other Mars Years. 3.) There is a decease in the water ice index around the
edge of the cap, particularly on the north side around Ls=92 in each Mars Year
(green arrows) and also on the south side of the ice mound (pink arrows).
~ 23 ~
Figure 6 – HiRISE images of Louth crater during spring (PSP_008530, Ls=75.1,
left) and mid-summer (ESP_037210, Ls=156, right). The close up images are
from PSP_008530 alone to eliminate seasonal differences. The close up images
are shown as boxes in the larger image, showing the rough stucco texture of the
exterior rim of the water ice mound. The images show the northern rim of the ice
mound contact with the regolith is clean and relatively distinct (blue arrows). The
southern rim of the ice mound is ragged and less distinct (red arrows). This is a
strong indication of southerly winds transporting water ice in the northerly
direction, and is in accord with the asymmetric offset of the ice mound within
Louth. The direction of this aeolian activity supports our hypothesis that winds
are responsible for the asymmetry of the ice mound in the crater. The green
arrows indicate a noticeable darkening of the regolith around the southern rim of
the crater in mid summer (right), relative to the spring image (left). This supports
~ 24 ~
our hypothesis that fine grained (bright) water ice is lying on the regolith during
spring, subliming during summer, and re-depositing onto the mound (see Figure
8). The linear features in the middle of the central ice mound (black lines) run
almost north-south, which is in accord with their interpretation as wind blown
sastrugi, which on Earth run parallel to the wind direction.
Figure 7 – Cumulative mass balance plot for a 1-D accumulation model for the
water ice mound at Louth (see model description in Bapst et al. (2017)). The
model assumes a constant windspeed of 3 m/s and is run for a range of albedo
of 0.35-0.43. CRISM observations constrain the timing in the transition from
accumulation to ablation (where zero is crossed in this plot) to Ls=83-100, which
is consistent with modeled albedo of 0.37-0.41 (solid lines).
~ 25 ~
Figure 8 – Diagram of H2O transfer process at Louth Crater during northern
spring and summer. (left) From Ls=0-60, Louth Crater is covered in CO2 ice,
shown in red. Water ice (blue) in the regolith is trapped. (center) At Ls=60, the
CO2 ice has sublimed and water ice in the regolith begins to sublime, and is
trapped on the nearby cold surface of the water ice mound, depositing a find
grained layer around the smooth edge of the ice mound. (right) From Ls=80-90,
no more water ice is available on the regolith around the ice mound, and the ice
around the boundary of the mound is fine grained and relatively smooth.
~ 26 ~
|
1503.07866 | 3 | 1503 | 2015-09-14T20:23:38 | Stellar and Planetary Properties of K2 Campaign 1 Candidates and Validation of 17 Planets, Including a Planet Receiving Earth-like Insolation | [
"astro-ph.EP",
"astro-ph.SR"
] | The extended Kepler mission, K2, is now providing photometry of new fields every three months in a search for transiting planets. In a recent study, Foreman-Mackey and collaborators presented a list of 36 planet candidates orbiting 31 stars in K2 Campaign 1. In this contribution, we present stellar and planetary properties for all systems. We combine ground-based seeing-limited survey data and adaptive optics imaging with an automated transit analysis scheme to validate 21 candidates as planets, 17 for the first time, and identify 6 candidates as likely false positives. Of particular interest is K2-18 (EPIC 201912552), a bright (K=8.9) M2.8 dwarf hosting a 2.23 \pm 0.25 R_Earth planet with T_eq = 272 \pm 15 K and an orbital period of 33 days. We also present two new open-source software packages which enable this analysis. The first, isochrones, is a flexible tool for fitting theoretical stellar models to observational data to determine stellar properties using a nested sampling scheme to capture the multimodal nature of the posterior distributions of the physical parameters of stars that may plausibly be evolved. The second is vespa, a new general-purpose procedure to calculate false positive probabilities and statistically validate transiting exoplanets. | astro-ph.EP | astro-ph |
Draft version Wednesday 16th September, 2015
Preprint typeset using LATEX style emulateapj v. 5/2/11
STELLAR AND PLANETARY PROPERTIES OF K2 CAMPAIGN 1 CANDIDATES AND VALIDATION OF 17
PLANETS, INCLUDING A PLANET RECEIVING EARTH-LIKE INSOLATION
Benjamin T. Montet1,2, Timothy D. Morton3, Daniel Foreman-Mackey4,5, John Asher Johnson2,
David W. Hogg4,5,6, Brendan P. Bowler1,7, David W. Latham2, Allyson Bieryla2, Andrew W. Mann8,9,
(Dated: Wednesday 16th September, 2015, 03:20)
Draft version Wednesday 16th September, 2015
ABSTRACT
The extended Kepler mission, K2, is now providing photometry of new fields every three months
in a search for transiting planets. In a recent study, Foreman-Mackey and collaborators presented
a list of 36 planet candidates orbiting 31 stars in K2 Campaign 1. In this contribution, we present
stellar and planetary properties for all systems. We combine ground-based seeing-limited survey data
and adaptive optics imaging with an automated transit analysis scheme to validate 21 candidates as
planets, 17 for the first time, and identify 6 candidates as likely false positives. Of particular interest
is K2-18 (EPIC 201912552), a bright (K =8.9) M2.8 dwarf hosting a 2.23 ± 0.25 R⊕ planet with
Teq = 272 ± 15 K and an orbital period of 33 days. We also present two new open-source software
packages which enable this analysis. The first, isochrones, is a flexible tool for fitting theoretical
stellar models to observational data to determine stellar properties using a nested sampling scheme to
capture the multimodal nature of the posterior distributions of the physical parameters of stars that
may plausibly be evolved. The second is vespa, a new general-purpose procedure to calculate false
positive probabilities and statistically validate transiting exoplanets.
Subject headings: catalogs — planetary systems — planets and satellites: detection — stars: funda-
mental parameters
1.
INTRODUCTION
The Kepler telescope (Borucki et al. 2010) has led to
a revolution in stellar and planetary astrophysics, with
7305 “objects of interest” and 4173 “planet candidates”
discovered to date (Borucki et al. 2011a,b; Batalha et al.
2013; Burke et al. 2014; Mullally et al. 2015; Rowe et al.
2015). The fidelity of this sample is high: most of these
candidates are truly planets (Morton & Johnson 2011;
Fressin et al. 2013; Désert et al. 2015). The mechani-
cal failure of two reaction wheels on the spacecraft led
to a repurposing of the spacecraft into the K2 mission,
in which the telescope points at fields near the ecliptic
plane for ∼ 75 days at a time (Howell et al. 2014). In
this observing strategy, two axes of motion of the space-
craft are controlled by the two remaining reaction wheels,
while the roll of the spacecraft is balanced with Solar ra-
diation pressure and quasiperiodic thruster firing. As a
result, the detector drifts relative to the sky at the rate
of ∼ 1(cid:48)(cid:48) hr−1, with rapid corrections due to thruster fires
[email protected]
1 Cahill Center for Astronomy and Astrophysics, California
Institute of Technology, Pasadena, CA, 91125, USA
2 Harvard-Smithsonian Center for Astrophysics, Cambridge,
3 Department of Astrophysics, Princeton University, Prince-
MA 02138, USA
ton, NJ, 08544, USA
4 Center for Cosmology and Particle Physics, Department of
Physics, New York University, 4 Washington Place, New York,
NY, 10003, USA
5 Center for Data Science, New York University, 726 Broad-
way, 7th Floor, New York, NY, 10003, USA
6 Max-Planck-Institut für Astronomie, Königstuhl 17, D-
69117 Heidelberg, Germany
7 Caltech Joint Center for Planetary Astronomy Fellow
8 Department of Astronomy, The University of Texas at
Austin, Austin, TX 78712, USA
9 Harlan J. Smith Fellow, The University of Texas at Austin
approximately once every six hours. Over the full dura-
tion of each campaign, the targets remain near the same
location on the detector but both the slow drift and the
corrections are observable by eye (Barentsen 2015).
K2 light curves produced with aperture photometry
contain substantial pointing-induced photometric varia-
tions caused by the star’s apparent motion over a poorly
defined flat field. Worse yet, these variations occur on
timescales similar to transit signals, potentially masking
the observational signature of a planet passing between
Kepler and its host star.
There has been considerable effort to recover these
planetary signals, and to date six planets have been con-
firmed orbiting three stars in the K2 data (Armstrong
et al. 2015; Crossfield et al. 2015; Vanderburg et al. 2015).
What is common to all of these methods are that removal
of systematics is considered a step to be undertaken be-
fore the search for planets. Under this strategy, it is
implicitly assumed that the systematics are removed per-
fectly, while retaining all of the astrophysical signal. Of
course, it is impossible to perfectly separate the astro-
physical and instrumental signal, and such a technique
is prone to either over-fitting, in which some of the as-
trophysical signal is also removed, or under-fitting, in
which some of the instrumental systematics remain. A
better strategy is to simultaneously fit both the signal
and the systematics, as is common practice in cosmology
and, increasingly, in radial velocity searches for plane-
tary systems (e.g. Ferreira & Jaffe 2000; Boisse et al.
2011; Haywood et al. 2014; Grunblatt et al. 2015).
Foreman-Mackey et al. (2015) simultaneously fit both
the systematics and potential planetary transit signals
in a search for transiting planets. They assume that
the dominant trends in the observed stellar light curves
2
are caused by spacecraft motion and are shared by
many stars. They then run principal component anal-
ysis (PCA) on all stars to measure the dominant modes,
modeling each star as a linear combination of 150 of
these “eigen light curves” and a transit signal. This
method enables fitting without over-fitting, and also per-
mits marginalization over uncertainties induced by the
systematic model. Therefore, any uncertainties in the
systematics can be propagated into uncertainties in de-
tected planet parameters, instead of assuming the sys-
tematics are understood perfectly. Using this technique,
Foreman-Mackey et al. (2015) detect 36 planet candi-
dates orbiting 31 stars in K2 Campaign 1 data.
In Foreman-Mackey et al. (2015), only transit prop-
erties are provided, not absolute parameters about the
planet or the star. Additionally, the authors follow the
convention of the Kepler team to include any transit
event as a candidate system rather than a false positive if
a secondary eclipse is not detected: there is no enforced
upper limit on the allowed planet radius. The authors
intentionally make no effort to separate true transiting
planets from astrophysical events that mimic the appear-
ance of transits, such as an eclipsing binary (EB) with
a high mass ratio, similar to the Kepler team’s list of
“objects of interest.”
In this paper, we present stellar and planetary pa-
rameters for each system. We also analyze the false
positive probability (FPP) of each system using vespa,
a new publicly available, general-purpose implementa-
tion of the Morton (2012) procedure to calculate FPPs
for transiting planets. Through this analysis, as well
as archival imaging, ground-based seeing-limited survey
data, and adaptive optics imaging, we are able to con-
firm 21 of these systems as transiting planets at the 99%
confidence level. Additionally, we identify six systems as
false positives.
This paper is organized as follows. In Section 2, we de-
velop stellar properties through photometric and spectro-
scopic data. In Section 3, we combine the derived stellar
properties with K2 data to infer planet candidate prop-
erties. In Section 4, we combine adaptive optics and ra-
dial velocity observations with both archival and modern
ground-based, seeing limited survey data and an analysis
of the transit parameters to calculate FPPs. In Section
5, we discuss potentially interesting systems, including a
mini-Neptune orbiting an M dwarf which receives a sim-
ilar insolation to the Earth. In Section 6, we summarize
and discuss our results.
2. STELLAR PROPERTIES
2.1. Photometry
With the exception of one star in our sample (K2-18),
we do not have spectroscopic data with which to char-
acterize the stellar properties. Additionally, there are
no measured parallaxes for any of these stars. Instead,
we rely on photometry. For each system, we query the
VizieR database of astronomical catalogs (Ochsenbein
et al. 2000). We record the B, V , g(cid:48), r(cid:48), and i(cid:48) mag-
nitudes and their uncertainties from the AAVSO Pho-
tometric All-sky Survey (APASS) DR6 (Henden & Mu-
nari 2014), as reported in the UCAC4 Catalog (Zacharias
et al. 2012). We also record the J, H, and K mag-
nitudes and their uncertainties as found in the 2MASS
All-sky Catalog of Point Sources (Cutri et al. 2003) and
the W 1 − W 3 WISE magnitudes and uncertainties from
the ALLWise Data Release (Cutri et al. 2013). For all
except two of our targets, the W 4 band is only an up-
per limit, and in the remaining two cases, the photomet-
ric uncertainity in W 4 is at least an order of magnitude
larger than those in W 1 − W 3, so we do not use W 4 for
any system. These data are reported in Table 1, and a
color-color diagram showing the r − J, J − K colors of
our candidates is included as Fig. 1.
Fig. 1.— Color-color diagram displaying r−J, J−K photometry
for targets observed by Kepler during the original mission (black),
with our K2 Campaign 1 planet candidates overlaid (red). Also
included is the location of the Sun (yellow) and host stars of pre-
viously confirmed K2 planets (blue). 90% of our candidates have
photometry consistent with later spectral types than the Sun.
2.2. Stellar Models
To convert the observed photometric data into physical
properties for each star, we used the new publicly avail-
able isochrones Python module,10 a general-purpose in-
terpolation tool for the fitting of stellar models to pho-
tometric or spectroscopic parameters (Morton 2015a).
This software does trilinear interpolation in mass–age–
[Fe/H] space for any given set of model grids, thus being
able to predict the value for any physical or photometric
property provided by the models at any values of mass,
age, and [Fe/H] within the boundaries of the grid.
This enables a set of observed properties ({xi, σi}), ei-
ther spectroscopic, photometric, or both, to define a like-
lihood function to be sampled:
(cid:88)
lnL(θ) ∝ − 1
2
(xi − Ii (θ))2
i
σ2
i
,
(1)
where Ii(θ) is the isochrone model prediction of property
i at the given parameters θ. If the observed properties in-
clude any apparent magnitudes, then θ includes distance
and extinction in addition to mass, age, and [Fe/H] .
In this work, we use grids from the Dartmouth Stellar
Evolution Database (Dotter et al. 2008) at Solar values
of [α/Fe] =0.0 and helium abundance Y = 0.2741, which
come packaged with the isochrones module. We then
infer the stellar parameters using MULTINEST (Feroz et al.
10 http://github.com/timothydmorton/isochrones
0.00.30.60.91.2J-K01234r-J2009), an implementation of a multimodal nested sam-
pling algorithm, for each host star conditioned on the
observed photometric properties as presented in Table
1. MULTINEST is designed to sample multimodal poste-
riors, where other samplers such as MCMC algorithms
often struggle. Given the multimodal nature of our pos-
teriors, this scheme is optimal for capturing parameter
space on the subgiant branch where these stars could re-
side. We include a prior on stellar metallicity represen-
tative of the observed metallicities of stars within 1 kpc
of the Sun, following the results of Hayden et al. (2015),
and a Salpeter-slope prior on mass up to the maximum
mass available in the model grids of 3.7 M(cid:12).
During the sampling process, we fit for Galactic ex-
tinction as one of our physical parameters. We include
the WISE bandpasses by applying the relative extinc-
tion values between SDSS, 2MASS, and WISE calcu-
lated by Davenport et al. (2014).
In each step of our
fitting process, we draw a value for AV , calculate the
expected extinction in all bandpasses AX assuming the
RV = 3.1 reddening law of (Fitzpatrick 1999), and then
measure the likelihood of our model stellar fit to the
observed apparent magnitudes. We apply a uniform
prior ranging from zero to a maximum extinction value
of 0.2 and marginalize over extinction in our final de-
termination of stellar parameters. The NASA/IPAC
Extragalactic Database, which reports the Schlafly &
Finkbeiner (2011) recalibration of the Schlegel et al.
(1998) extinction map as measured by COBE/DIRBE
and IRAS/ISSA, suggests that typical AV extinction val-
ues to the edge of the galaxy at this high Galactic latitude
are ∼0.1 mag, so our upper limit appears to be justified.
Such a scheme enables us to infer the statistical uncer-
tainties on the mass, radius, and effective temperature.
However, we are subject to biases induced by system-
atics in the models themselves. There is some evidence
that the Dartmouth models may under-predict radii of
M dwarfs by ∼ 15% when compared to other methods
(Newton et al. 2015; Montet et al. 2015). Such an ef-
fect may be the result of the Dartmouth model reliance
on BT-Settl atmospheres, which are based on incomplete
molecular line lists and have been shown to predict near-
IR colors that are too blue (Thompson et al. 2014).
As our stellar results are model-dependent, we caution
users who intend to use these parameters for other works,
such as exoplanet population studies. When available,
stellar parameters inferred through other techniques such
as asteroseismology or spectroscopy should supersede
these values. We note the observed photometric param-
eters are consistent with spectroscopically derived pa-
rameters for stars with published spectra, and consistent
with typical model-dependent uncertainties from photo-
metric data (e.g. Huber et al. 2014). We provide full
samples of our posteriors on the physical parameters for
each star.11
Bastien et al. (2014) use the “granulation flicker” in the
Kepler light curves to suggest that approximately 50%
of planet host stars have evolved off the main sequence
onto the subgiant branch, so that both the host stars and
their planets are larger than previously reported. Simi-
larly, in K2 Campaign 1 we may expect to find evolved
stars in a sample of planet candidates, although we may
3
expect the effect to be lessened due to the high Galac-
tic latitude of Campaign 1. Indeed, we find this to be
the case. Two stars, EPIC 201257461 and 201649426 are
definitively evolved stars, with inferred masses less than
2 M(cid:12)but radii above 8 R(cid:12). For approximately one third
of the others, we find the stellar radius posterior dis-
tribution to be bimodal, with both main sequence and
subgiant models of the stars being consistent with the
photometric data. This number is consistent with our
expectations of the number of subgiant contaminants in
the Campaign 1 field (K. Stassun 2015, private commu-
nication). Future observations to measure the parallaxes
of these stars, such as with Gaia, will be helpful in dif-
ferentiating between these two models to determine more
precisely the stellar, and thus the planetary, radii.
2.3. SuperNova Integral Field Spectrograph (SNIFS)
and SpeX Spectroscopy
A near-infrared spectrum of K2-18 was obtained using
the upgraded SpeX (uSpeX) spectrograph (Rayner et al.
2003) on the NASA Infrared Telescope Facility (IRTF)
on 2015 January 29 (UT). SpeX observations were taken
using the short cross-dispersed mode and the 0.3×15(cid:48)(cid:48)
slit, which provides simultaneous coverage from 0.7 to
2.5µm at R (cid:39) 2000. The target was observed at two
positions along the slit to subsequently subtract the sky
background. Eight spectra were taken following this pat-
tern, which provided a final signal-to-noise ratio (S/N)
of > 150 per resolving element. The spectrum was flat
fielded, extracted, wavelength calibrated, and stacked
using the Spextool package (Cushing et al. 2004). An
A0V-type star was observed immediately after the tar-
get, which was used to create a telluric correction using
the xtellcor package (Vacca et al. 2003).
An optical spectrum was obtained using SNIFS (Alder-
ing et al. 2002; Lantz et al. 2004) on the University of
Hawai’i 2.2m telescope on the night of 2015 January 30.
SNIFS provides simultaneous coverage from 3200–9700Å
at a resolution of (cid:39) 1000. Final S/N of the spectrum
was > 100 per resolving element in the red (∼ 6000Å).
Details of the SNIFS reduction, including dark, bias, and
flat-field corrections, cleaning the data of bad pixels and
cosmic rays, and extraction of the one-dimensional spec-
trum are described in Bacon et al. (2001) and Aldering
et al. (2006). Flux calibration was performed using a
separate pipeline described in Mann et al. (2015).
Teff was calculated by comparing our optical spectra
with the CFIST suite12 of the BT-SETTL version of
the PHOENIX atmosphere models (Allard et al. 2013),
which gave a temperature of 3503 ± 60 K. More details
of this procedure are given in Mann et al. (2013b) and
Gaidos et al. (2014). This method was used because it
is known to accurately reproduce empirical Teff values
from long-baseline optical interferometry Boyajian et al.
(2012).
Metallcity was determined using the procedures from
Mann et al. (2013a), in which the authors provide em-
pirical relations between atomic features and M dwarf
metallicity, calibrated using wide binaries. We adopted
the weighted mean of the H− and K−band calibrations,
which yielded a metallicity of 0.09±0.09.
11 http://www.astro.princeton.edu/~tdm/k2/
12 http://phoenix.ens-lyon.fr/Grids/BT-Settl/CIFIST2011/
4
We combined the derived Teff and [Fe/H] values with
the empirical Teff-[Fe/H]-R∗ relation from Mann et al.
(2015) to compute a radius. Accounting for measurement
and calibration errors in [Fe/H] and Teff we calculated a
radius 0.394±0.038R(cid:12). We use these parameters instead
of the derived photometric properties for this target, al-
though we note the two are consistent at the 1σ level.
The full list of stellar parameters adopted in this paper
is included in Table 2.
3. PLANET PROPERTIES
In Foreman-Mackey et al. (2015), only parameters di-
rectly observable from the K2 light curve itself were re-
ported:
the period, time of transit center, and tran-
sit depth. With stellar properties now in hand, we
can convert these observational results into fundamen-
tal parameters of each planet candidate. For each can-
didate, we fit the light curve using a physical transit
model (Mandel & Agol 2002; Kipping 2010) simultane-
ously with a systematics model similar to the one de-
scribed by Foreman-Mackey et al. (2015). We use emcee
(Foreman-Mackey et al. 2013), an implementation of the
affine-invariant ensemble sampler of Goodman & Weare
(2010) to sample from the posterior probability distribu-
tion for the stellar—limb darkening coefficients, mass, ra-
dius, and effective temperature—and planetary—radius,
period, phase, impact parameter, eccentricity, and argu-
ment of periapsis—parameters, conditioned on the light
curve and the measured stellar properties.
Following Foreman-Mackey et al. (2015), the likelihood
function that we use is marginalized over the weights of
the “eigen light curves” in the linear systematics model.
Unlike Foreman-Mackey et al. (2015), we include an em-
pirical Gaussian prior on the weights determined by ro-
bustly computing the distribution of weights across the
full set of Campaign 1 light curves. This prior miti-
gates the incorrect detection of false signals induced by
stellar variability—as discussed below in Section 5.2—so
we exclude these candidates (EPIC 201929294 and EPIC
201555883) from the tables of results.
In this analysis, we assume the dilution caused by ad-
ditional stars contributing flux into the aperture is neg-
ligible for nearly all systems. Given the location of the
Campaign 1 field at a high Galactic latitude, we expect
low contamination by background giants. Nevertheless,
this assumption may not be valid for all systems. Any
contamination unaccounted for, as may happen if any of
these stars are actually unresolved binaries, would cause
us to underestimate the radii of any planets we detect.
Therefore, high-contrast adaptive optics imaging of any
systems should be obtained before these planets are used
in population inference studies. The planet parameters
measured by this analysis are listed in Table 3.
4. FALSE POSITIVE ANALYSIS
There are many scenarios which can cause an astro-
physical false positive, where an EB star masquerades
as a transiting planet. The most common scenarios
are if (a) it is a highly grazing eclipse, or (b) the bi-
nary system shares a photometric aperture with a sig-
nificantly brighter star, resulting in a diluted eclipse
depth. When possible, such astrophyscial false positive
scenarios are traditionally ruled out by detailed follow-
up observations, often a combination of high-resolution
imaging and radial-velocity measurements. However,
the Kepler mission, with its thousands of planet candi-
dates around mostly faint stars, necessitated a paradigm
shift—a move toward probabilistic interpretation of tran-
sit signals, rather than comprehensive follow-up of each
individual candidate (Morton & Johnson 2011).
Morton (2012) presented an automated method to cal-
culate the probability that a planet candidate might be
caused by an astrophysical false positive. This method
uses Galactic population simulations to determine the
distributions of possible false positive scenarios, compar-
ing the typical light curve shape of each to the data. It
then combines this information with observationally mo-
tivated prior assumptions about the populations of field
stars, the properties of multiple star systems, and the
occurrence rate of planets as determined from Kepler
(Fressin et al. 2013), in order to determine the proba-
bility that the observed signal may be a false positive.
Similar in spirit to other published methods of proba-
bilistic validation, such as BLENDER (Torres et al. 2011)
and PASTIS (Díaz et al. 2014), it has the advantage of
being computationally less demanding and fully auto-
mated, and thus easily applied in batch to a large number
of candidates.
In this work, we use vespa13 (Morton 2015b), a new
publicly available, general-purpose implementation of
the Morton (2012) procedure, to calculate FPPs for each
of these K2 candidates. The following constraints on
false positive scenarios are imposed:
• A chance-aligned EB system may reside anywhere
inside or within one pixel of the photometric aper-
ture of the target star. In creating a light curve for
each star, we define photometric apertures rang-
ing from 10 to 20 arcsec for each star, as defined
in Table 4. Given the 6 arcsec point spread func-
tion (PSF) of the Kepler telescope, we allow for
the possibility that companions falling just outside
of our aperture (within one pixel) may contribute
to the light curve, possibly causing a false positive
event. The search for such companions is discussed
in §4.3.
• The maximum allowed depth of a potential sec-
ondary eclipse event is the most significantly de-
tected signal at the same period of the planet can-
didate, once the primary transit is masked out (dis-
cussed in §4.1). vespa does not allow for the pos-
sibility of secondary eclipses larger than those ob-
served in the K2 light curve for each star.
• Blended stars must be allowed by the available
adaptive optics and archival
imaging data (dis-
cussed in detail in §4.2 and 4.3). vespa only con-
siders stars below the detection threshold for the
AO imaging, which is a position-dependent value
following a calculated contrast curve for each star.
Each of these scenarios is an astrophysical eclipse,
caused by one object passing in front of another, blocking
some fraction of the total light. The calculations here do
not include the possibility that each signal is caused by
13 http://github.com/timothydmorton/vespa
an instrumental artifact in the data or some other astro-
physical event, such as stellar activity, masquerading as
planet transits.
Table 5 summarizes the results of these calculations,
presenting the relative probability for each candidate
to be caused by any of three false positive scenarios:
an undiluted EB, a hierarchical triple eclipsing binary
(HEB), and a chance-aligned background(/foreground)
eclipsing binary (BEB).
Six of the presented candidates have FPP >90%; these
are considered to be likely false positives. On the other
hand, 24 candidates have FPP < 1%. Three of the tran-
sit signals might plausibly be caused by contamination
by detected stellar companions within the photometric
apertures (see §4.3), so we keep these as candidates.
This leaves 21 candidates that we statistically validate
as planets, including four that have been previously iden-
tified in the literature (Armstrong et al. 2015; Crossfield
et al. 2015). So in total, of the 36 candidates, 21 are
secure planets, 17 of which we validate here for the first
time.
We emphasize that the majority of these validations
rely solely on the transit photometry and SDSS data,
with follow-up imaging only obtained for seven of the
31 targets. This demonstrates the utility of the vespa
tool, which will be crucial to interpreting future candi-
dates detected by K2, TESS, and PLATO and priori-
tizing follow-up observing efforts. We show the transit
signals in Figure 2.
4.1. Secondary Eclipse Observations
One of the definitive signatures of a false positive bi-
nary star system masquerading as a transiting planet is
the presence of a secondary eclipse. While a nondetec-
tion of a secondary does not exclude the possibility of a
binary system (the orbit may be eccentric, or the com-
panion too faint for a secondary eclipse to be detectable
in the noise), such a nondectection reduces the probabil-
ity of each of the EB false positive scenarios.
To attempt to eliminate each EB scenario, we first
search each K2 light curve to determine which secondary
eclipse signals are not allowed by the data. We mask the
transit signal of the planet in question and search for the
most significant signal at the same period. Such a scheme
does not assume circular orbits: we return the most sig-
nificant signal at any phase, not only at the midpoint
between consecutive transits.
We report these maximum allowable secondary eclipse
depths in Table 5. These values are used by vespa as
limits on the allowable secondary eclipse. Any models
that cause a larger event, such as a background EB con-
sisting of two equal-mass stars in a circular orbit, can be
excluded by the data. We note that with the exception
of K2-19c, all systems with a maximum eclipse depth
of at least one part per thousand have FPPs of 0.866
or larger. The exception, K2-19, is a two-planet system
with the two planets near a 3:2 period commensurability,
so in this case the “secondary” is actually the transits of
the other planet.
4.2. Adaptive Optics Imaging
We obtained high resolution images of seven stars
with the Palomar High Angular Resolution Observer
5
(PHARO) infrared detector (Hayward et al. 2001) be-
hind the PALM3000 adaptive optics system (Dekany
et al. 2013) at the Palomar 5.1 m Hale telescope on the
nights of 2015 February 3 and 4 UT. Sky conditions were
mostly clear with light cirrus and ≈1.(cid:48)(cid:48)0–1.(cid:48)(cid:48)3 seeing on
both nights. We used the smallest plate scale of 25 mas
pixel−1 which resulted in a field of view of 25.(cid:48)(cid:48)6×25.(cid:48)(cid:48)6
across the 10242 pixel2 array. All observations were ob-
tained with the 32x pupil sampling mode, resulting in
Strehl ratios of ≈20–30% in KS for our V =11–13 mag
targets as measured by the Strehl monitor at the tele-
scope in real time. We obtained unsaturated dithered
frames of each target in KS-band with typical integra-
tion times of 2–10 s. Except for EPIC 201828749 and
EPIC 201546283, which had nearby candidate binary
companions, we also acquired deep saturated images (5–
10 frames at 60 s each) to search for fainter companions.
Images were registered and contrast curves were gen-
erated following Bowler et al. (2015). For the saturated
data, the star’s position in each image was found by
masking the saturated region and fitting a 2D bivari-
ate Gaussian to the PSF wings. Contrast curves for the
median-combined image are calibrated using the unsat-
urated frames. The typical sensitivity is 6.5–7.5 mag
at 1(cid:48)(cid:48). The images were astrometrically calibrated using
dithered observations of the Trapezium cluster centered
on θ1 Ori C taken on 2015 February 3 UT. Based on the
reference astrometry for pairs of stars in the field from
McCaughrean & Stauffer (1994), we measure a plate
scale of 25.2 ± 0.4 mas pixel−1 and north orientation of –
0.◦2 ± 0.◦3. Since this latter value is consistent with being
aligned with the detector columns, we adopt a value of
0.◦0 ± 0.◦3 for this work. Relative photometry of nearby
stars is carried out using aperture photometry with an
aperture radius of 12 pixels (0.(cid:48)(cid:48)3). For EPIC 201828749,
we also acquired J- and H-band images. Astrometry and
photometry is derived separately for each image, and the
mean and standard deviation of these measurements is
adopted for our final values listed in Table 4.
Images for all systems AO data was obtained for is
shown in Figure 3, while contrast curves showing the 5σ
limits for detection as a function of orbital separation are
given in Figure 4.
4.3. Known Background Stars
The PHARO AO system has a field of view of 25 arcsec.
Each K2 pixel is a square, 3.(cid:48)(cid:48)98 on a side. A background
EB within a few K2 pixels of our target stars could mimic
a transit signal inside our aperture while evading detec-
tion by PHARO. Such wide EBs should appear in seeing-
limited ground-based surveys.
To investigate the possibility that such wide compan-
ions exist, we query the ninth data release of the Sloan
Digital Sky Survey (SDSS DR9, Ahn et al. 2012). For
each target, from the depth of the observed transit we
determine how bright a background object must be to
cause the event if the background object were an equal
mass totally eclipsing binary. We then search for all stars
within 25(cid:48)(cid:48) that are within this brightness limit relative
to the candidate host star. All apertures we use in our
K2 analysis are smaller than 20(cid:48)(cid:48) so this search should
encompass the region where possible background con-
taminants could reside. Of the 31 stars in our sample,
eleven have such a companion, plus one detected in AO
6
Fig. 2.— Phase-folded K2 photometry for all planet candidates analyzed in this paper. Each is the product of a fiducial noise model, in
which the median systematic has been removed for illustrative purposes. The systems which we validate as transiting planets are labeled
in blue. The systems which we confirm as false positive events are labeled in red. The systems which we leave as candidates are labeled in
black. Red curves outline the median transit model for each candidate system.
imaging.
dates in Table 5, rather than planets. Further updates to
the vespa code will allow consideration of “specific” false
positive scenarios; that is, scenarios that correspond to
actually detected stars such as these, rather than hypo-
thetical background or bound companions.
The candidates with identified companions that we
judge to not be plausible sources of potential contami-
nation are the following:
• K2-13b (201629650.01)— The companion to this
star is 17.(cid:48)(cid:48)3 from the EPIC target. As this is out-
side the aperture (radius 15.(cid:48)(cid:48)9) and the background
star is not particularly bright, we rule out contam-
ination for this system.
• 201702477.01— The companion to this star is
12.(cid:48)(cid:48)15 from the EPIC target, and the aperture size
is 10.(cid:48)(cid:48)0. In addition, the maximum depth in this
system is almost identical to the transit depth. For
these two reasons we rule out contamination in this
case.
Unlike the original Kepler field, the field for K2 Cam-
paign 1 is well out of the Galactic plane, so the rate of
giant, distant background stars is significantly lower. We
include all potential contaminants in Table 4. We vali-
date or eliminate each of these as a possibility based on
the transit shape. For example, the events near EPIC
201546283 could only be caused by a background binary
if the background object was a completely eclipsing sys-
tem (so that the eclipse depth was 50%). In this case,
the transit would be V-shaped. Since it is not, the back-
ground object likely does not cause the transit event.
In Table 4, the “maximum depth” column represents
the maximum observed “transit” depth if the transit were
actually caused by a total eclipse of the hypothetical
background binary system, inducing a 50% flux decre-
ment in the background star’s apparent brightness.
The photometric apertures used to detect these can-
didates range in radius from 10.(cid:48)(cid:48)0 to 19.(cid:48)(cid:48)9. In order to
be a plausible contaminant, any companion star must
be either within this aperture or just outside but bright
enough for signficant flux to leak in. Evaluating each of
the systems listed in Table 4, we judge that we cannot yet
rule out contamination as a potential source of the transit
signal for four candidates: 201295312.01, 201403446.01,
201546283.01, and 201828749.01. Despite receiving low
FPP scores from vespa, we list these systems as candi-
SDSS is 95% complete at r = 22.2 mag and the tele-
scope has a PSF of 1.(cid:48)(cid:48)4. For the purposes of the vespa
calculation, we thus treat nondetection in SDSS data as
providing a contrast curve at wide separations down to
a limiting magnitude of r = 22.2 mag.
4.4. Archival Imaging
-3.0-1.01.03.0201208431.01-2.5-1.5-0.50.51.5201257461.01-0.8-0.40.00.40.8201295312.01-3.0-1.01.03.0201338508.01-3.0-1.01.03.0201338508.02-2.0-1.00.01.0201367065.01-1.5-1.0-0.50.00.5201367065.02-1.0-0.6-0.20.20.6201384232.01-1.5-0.50.51.5201393098.01-1.0-0.6-0.20.20.6201403446.01-2.5-1.5-0.50.51.5201445392.01-2.5-1.5-0.50.51.5201445392.02-5.0-3.0-1.01.03.0201465501.01-10.0-6.0-2.02.0201505350.01-10.0-6.0-2.02.0201505350.02-3.5-2.5-1.5-0.50.5201546283.01-1.5-0.50.51.5201549860.01-30.0-20.0-10.00.010.0201565013.01-100.0-60.0-20.020.0201569483.01-2.5-1.5-0.50.5201577035.01-1.5-0.50.5201596316.01-0.8-0.40.00.40.8201613023.01-2.0-1.00.01.02.0201617985.01-1.0-0.50.00.51.0201629650.01-15.0-10.0-5.00.05.0201635569.01-120.0-80.0-40.00.0201649426.01-7.0-5.0-3.0-1.01.0201702477.01-3.0-1.01.03.0201736247.01-2.0-1.00.01.02.0201754305.01-2.0-1.00.01.02.0201754305.02-0.4-0.20.00.20.4 Time from Transit Center (Days)-40.0-30.0-20.0-10.00.0 Relative Flux (ppt)201779067.01-0.4-0.20.00.20.4-1.5-0.50.5201828749.01-0.4-0.20.00.20.4-1.5-0.50.51.5201855371.01-0.4-0.20.00.20.4-4.0-2.00.02.0201912552.017
Fig. 3.— Adaptive optics images for the seven stars observed with high-contrast imaging. The main frame for each single system shows
the deep, saturated image. The inset for each single system shows a shallower, unsaturated image to better identify companions at close
projected orbital separations. For the two systems with imaged companions, EPIC 201546283 and EPIC 201828749, only unsaturated
frames are collected. The pixel scale is 0.(cid:48)(cid:48)0252 per pixel. Each subplot is a square 400 pixels on a side and each inset is a square 100 pixels
on a side. All subplots, including insets, are plotted on the same scale.
larger than 50 mas yr−1, so they have moved across the
sky by (cid:38) 2.(cid:48)(cid:48)5 since they were imaged during the first
Palomar Observatory Sky Survey (POSS) in the 1950s.
To rule out background companions, we download data
from the POSS I and II surveys, which imaged these
targets in 1952-1955 and 1989-1998, respectively. We
also download data from the Sloan Digital Sky Survey,
which imaged these fields between 2000 and 2009. As
shown in Figure 5, we do not detect any background
targets at the present-day location of any of these stars
For the stars with AO nondetections, there is still the
possibility that a background binary could be positioned
directly behind the target star, evading detection. The
probability is small, given the 0.(cid:48)(cid:48)1 diffraction limit of the
Hale Telescope at 2 µm, but nonzero. While the vespa
calculations quantify this probability for this to occur,
we can also rule out the possibility of such chance align-
ments, down to a certain contrast, with archival imaging
data.
Five of the stars in our sample have proper motions
2012953122014034462015462832015770352016130232018287495''2019125528
5. POTENTIALLY INTERESTING SYSTEMS
5.1. A Mini-Neptune with Earthlike Insolation
The planet orbiting K2-18 may be an interesting target
for atmospheric studies of transiting exoplanets.
By combining archival and modern seeing-limited data
with adaptive optics imaging, we can exclude the possi-
bility these transit events are caused by a background
EB. The apparent transits must be caused by an ob-
ject co-moving with K2-18; radial velocities eliminate the
possibility the companion is nonplanetary. Therefore, we
confirm the planetary nature of this system.
This star is an M2.8 dwarf at a distance of 34±4 pc. Of
our planet candidate hosts, only K2-3 (originally discov-
ered by Crossfield et al. 2015) is brighter in K-band. This
star is only 0.1 mag fainter in K than GJ 1214 (Charbon-
neau et al. 2009). Due to the relative brightness of the
host star, this target is likely to become a prime target
for atmospheric characterization studies and is ideal as a
target for future space-based missions such as JWST.
The planet is slightly smaller than GJ 1214b, but un-
like that planet, K2-18b is not highly irradiated. Instead,
it is at a reduced semimajor axis a/R(cid:63)= 83.8 ± 9.0. Its
equilibrium temperature is then, assuming zero albedo,
Teq = 272 ± 15 K, meaning its bulk insolation is 94 ± 21
percent that of the Earth’s. Although the planet is likely
too large to be rocky (Rogers 2015), its atmosphere is
likely to be the focus of many future observations, pro-
viding a cool analogue to the highly irradiated planets of
a similar size found by Kepler.
5.2. Other Sources of False Positives
The method of Foreman-Mackey et al. (2015) assumes
that all variability in the light curves are caused by ei-
ther the motion of K2, in which case the variability is
shared by all stars, or transits of planets, in which case
the variability is intrinsic to only one star. This assump-
tion breaks down for extremely spotted stars where the
astrophysical variability is larger than the instrumental
magnitude.
In that regime, the starspot modulations
can be incorrectly fit by the systematic model, causing
spurious transits to appear. This appears to be the case
with EPIC 201929294, which has coherent starspots that
appear to have the same rotation period as the transit
period reported previously. Because the starspots are so
periodic and coherent, these spurious transits were falsely
identified as a planet candidate; we consider that system
a false positive in this work.
object
orbiting EPIC
201555883 has a period, time of transit, and tran-
sit duration consistent with EPIC 201569483.
Such
effects are not uncommon in Kepler data. Coughlin
et al. (2014) identify 685 KOIs as false positives and
outline four physical reasons why these anomalies may
occur. EPIC 201555883 is a unique case in that it does
not appear to fall under any of these cases.
It falls
on module 23, while EPIC 201569483 is on module 8,
neither 180◦ away from nor on the same column as
this candidate. Moreover, there is not any evidence of
a mechanism that could cause a third star to induce
both the appearance of a 7% eclipse on one module and
an additional anomalous transit event on a different
module. Instead, this candidate could be a false positive
caused by a different systematic mechanism.
candidate
possibly
The
Fig. 4.— 5σ contrast curves for all systems with AO nondetec-
tions. For all systems, we can exclude the possibility that a com-
panion at a given ∆KS exists. From our known transit depths, we
can then rule out significant parameter space in which an eclipsing
binary could reside and mimic a transit signal.
in any of these images.
For these targets, we can extend our contrast curves
to zero present-day orbital separation and rule out the
possibility that these transit events are caused by a back-
ground EB. By combining present-day seeing-limited
photometric survey data, adaptive optics imaging, and
archival photometry, the only stellar companions we
would not detect would be those that are gravitationally
bound to the target star and positioned in their orbits so
that their projected separation is smaller than the diffrac-
tion limit of the Hale Telescope. Such an alignment
would require the orbital inclination of the binary to be
nearly 90◦ and the phase + θ ≈ π/2 or 3π/2. While
we cannot fully rule out this possibility, the vespa cal-
culations confirm that its probability is negligibly small.
4.5. Tillinghast Reflector Echelle Spectrograph (TRES)
Radial Velocities
We observed K2-18 on 2015 February 04 and 25 UT
with TRES on the 1.5 m Tillinghast Reflector at the Fred
L. Whipple Observatory. These dates were chosen to be
near the times of largest RV variations, corresponding
to phases of 0.72 and 0.32 relative to the time of tran-
sit. The spectra were taken with a resolving power of
R = 44, 000 and integration times ranging from 2800 to
3600 s, resulting in S/N between 17 and 29 per resolution
element.
The spectra were extracted as described in Buchhave
et al. (2010). The relative RVs were derived by cross-
correlating the spectra against the strongest observed
spectrum (in this case, the first) over the wavelength
range 4700 - 6800 Å. We selected 19 echelle orders in
the analysis, being careful to reject orders with telluric
absorption lines, fringing in the far red and those with
very low SNR in the blue.
The two observed spectra have RVs that differ by 47±
42 m s−1. If the RVs were caused by a stellar companion,
the RV shift between these observations would be on the
order of km s−1. Therefore, we can rule out any stellar-
mass companions that would be able to create this transit
signal.
110Separation (")108642Contrast ((cid:54)KS)EPIC 201295312EPIC 201403446EPIC 201577035EPIC 201613023EPIC 2019125529
Fig. 5.— Archival maging for the five highest proper motion targets in our sample. In all cases, there are no background objects directly
behind the present day location of the target (red circle) that could be missed by the AO observations. Modern SDSS imaging can also
rule out wide companions that may have been missed at wide separations, beyond the AO field of view, such as the companion which can
be seen in the images of K2-10. All figures are aligned such that north is up and east to the left. All subplots are on the same scale.
POSS I, 1955POSS II, 1989201367065SDSS, 2000POSS I, 1952POSS II, 1998201465501SDSS, 2009POSS I, 1955POSS II, 1989201577035SDSS, 2000POSS I, 1952POSS II, 1998201613023SDSS, 2000POSS I, 1955201912552POSS II, 1989SDSS, 200210''10
Foreman-Mackey et al. (2015) modeled the systematic
effects in the K2 light curves using a linear combination
of “eigen light curves” generated empirically by running a
principle component analysis on the light curves of every
star. This means that the training set includes the light
curves for variable stars, EBs, and even transiting plan-
ets. Again, this star has significant variability caused by
starspots. In this case, the fitting procedure tries to ac-
count for stellar variability using the eigen light curves.
This overfit gives undue weight to eigen light curves that
include the transits of EPIC 201569483, causing this spu-
rious transit to occur. Again, we consider this system to
be a false positive. As stated in Section 3, by including
an empirical Gaussian prior on the weights for the eigen
light curves in the linear systematics model, the signals
observed on EPIC 201555883 and 201929294 are miti-
gated, suggesting such a scheme should be employed in
searching for planet candidates in future campaigns.
The problem of over-fitting stellar variability using
eigen light curves can also be solved by adding a stel-
lar activity model to our fitting procedure. In this case,
the spacecraft motion could be fit simultaneously with a
model of starspot modulation, astroseismic oscillations,
and planet transits. Such a model is currently under
development (Angus et al. 2015, in preparation)
5.3. Multiple Planet Systems
Five of the systems reported by Foreman-Mackey et al.
(2015) have more than one transiting candidate. One of
these is K2-3, a three-planet system originally announced
by Crossfield et al. (2015). Another of these is K2-19
(Armstrong et al. 2015), a two-planet system with the
orbital periods of the two planets near a 3:2 period com-
mensurability. The remaining three are all representa-
tive of the multiple-planet systems observed by Kepler
(Lissauer et al. 2011; Fabrycky et al. 2014). Two of the
systems are near a period commensurability and all three
consist of mini-Neptune sized planets.
We do not detect any significant transit timing varia-
tions (TTVs) in any of these systems from the K2 data
alone. K2-5 would be expected to have a TTV period
of 117 days, but is likely too far from commensurability
to have an observable TTV signal. K2-8 is expected to
have a TTV period of 234 days, so this system may be a
candidate for additional follow-up to constrain the sys-
tem masses dynamically. The transiting planets orbiting
K2-16 are near a 5:2 period commensurability. There
is no evidence from Kepler of an abundance of planets
near this period ratio, and so this may be coincidence.
Follow-up observations may be warranted to search for
an additional planet in this system forming a resonant
chain, similar to those observed around other stars (e.g.
Swift et al. 2013; Campante et al. 2015).
5.4. Systems Orbiting Bright Stars
One of the primary goals of K2 is the detection of tran-
siting planets around bright stars that can be followed up
from the ground or with future space-based observatories
such as JWST (Howell et al. 2014). Of our sample, two
systems orbit stars with K < 9 mag: K2-3 (Crossfield
et al. 2015) and K2-18. An additional planet candidate
may orbit EPIC 201828749, a star with K = 9.93 ± 0.03
mag. These targets are ideal for ground-based followup
and may be useful targets for Spitzer and JWST to probe
planetary atmospheres.
6. RESULTS AND DISCUSSION
We have presented stellar parameters for all planet
candidates systems identified by Foreman-Mackey et al.
(2015). We statistically validate 21 of the 36 candidates
as bona fide planets, and we identify 6 as false positives,
including two systematic false alarms. Of the planets, 4
have been previously validated in other works, while 17
are validated here for the first time. The systems not
validated as planets or false positives remain as planet
candidates.
Enabling much of this analysis are two new open-source
Python packages: isochrones14, which we use to infer
posteriors on physical stellar properties based on fitting
theoretical stellar models to observed data; and vespa15,
a new implementation of the Morton (2012) transit false
positive analysis scheme. Both of these packages will
continue to be useful in future analysis of transit candi-
dates where comprehensive follow-up observations may
be unavailable.
The isochrones package uses the nested sampling
scheme MULTINEST to capture the true multimodal na-
ture of the posteriors. Using an MCMC algorithm in-
stead can cause only one peak in the posterior distri-
bution to be sampled.
If the photometry is consistent
with both a star on the main sequence and the sub-
giant branch, an MCMC technique could cause one of
these peaks (likely the subgiant possibility) to be missed,
leading to an underestimation in the likelihood of sub-
giant stars and and underestimation of the uncertainties
of both the stellar and planetary parameters.
With the exception of one object, all of the stellar pa-
rameters are derived from comparing photometric obser-
vations to the Dartmouth stellar evolution models. As a
result, both the stellar and planet parameters are subject
to systematic biases induced by discrepancies between
the models and reality.
The planets we confirm in this paper, like the planets
found in the original Kepler mission, span a wide range
of parameter space. They are at distances ranging from
34 to 700 pc, have radii ranging from 1.3 to 5.3 R⊕,
and orbit with periods ranging from 5.0 to 50.3 days.
Like the original mission, we find significantly more small
planets than large planets, as expected from the radius
distributions measured from Kepler (Howard et al. 2012;
Fressin et al. 2013; Morton & Swift 2014).
Unlike the original mission, however, we find that
nearly all of our confirmed planets are around stars less
massive than the Sun. This difference is a result of both
the Campaign 1 field and the target selection process.
Campaign 1 is at a significantly higher Galactic lati-
tude than the original Kepler mission, meaning there is a
much lower number density of targets at large distances.
As massive stars at kiloparsec distances are relatively
less likely to exist in Campaign 1 than near the Galac-
tic plane, the pool of targets that could be selected for
Campaign 1 contains a larger fraction of subsolar stars.
Low-mass stars, particularly M dwarfs, are also a spe-
cific focus of the K2 mission. One of the primary goals
14 http://github.com/timothydmorton/isochrones
15 http://github.com/timothydmorton/vespa
of the Kepler mission was to “determine the abundance
of terrestrial and larger planets in or near the habitable
zone of a wide variety of spectral types of stars” (Batalha
et al. 2013) However, ∼ 70% of Kepler’s target stars had
masses within 20% of the Sun’s, while 70% of the stars
in the Galaxy have less than 50% the mass of the Sun
(Brown et al. 2011). K2 will fulfill the promise of Kepler,
with the goal of providing a yield of small planets around
bright, small stars to facilitate follow-up measurements
(Howell et al. 2014). This is clear from the K2 target
selection process, with thousands of K and M dwarfs be-
ing selected in each campaign. Based on these plans, we
expect that K2 will detect hundreds of planets during
its lifetime, with the majority being mini-Neptunes and
super-Earths around stars less massive than the Sun.
We thank Roberto Sanchis-Ojeda (Berkeley), Dan Hu-
ber (Sydney) and Jeff Coughlin (SETI) for conversa-
tions and suggestions which improved the quality of this
manuscript. We also thank Keivan Stassun (Vanderbilt)
for his insights into stellar parameters and the rate of
subgiant contamination for both Kepler and K2, which
significantly improved this work. We thank the anony-
mous referee for their comments and suggestions.
We are grateful to the entire Kepler team, past and
present. Their tireless efforts were all essential to the
tremendous success of the mission and the successes of
K2 present and future.
Some of the data presented in this paper were ob-
tained from the Mikulski Archive for Space Telescopes
(MAST). STScI is operated by the Association of Uni-
versities for Research in Astronomy, Inc., under NASA
contract NAS5–26555. for MAST for non–HST data is
provided by the NASA Office of Space Science via grant
NNX13AC07G and by other grants and contracts.
This paper includes data collected by the Kepler mis-
sion. Funding for the Kepler mission is provided by the
NASA Science Mission directorate.
11
This paper includes data collected by the Sloan Digital
Sky Survey. Funding for SDSS-III has been provided by
the Alfred P. Sloan Foundation, the Participating Insti-
tutions, the National Science Foundation, and the U.S.
Department of Energy Office of Science. The SDSS-III
web site is http://www.sdss3.org/. SDSS-III is managed
by the Astrophysical Research Consortium for the Partic-
ipating Institutions of the SDSS-III Collaboration includ-
ing the University of Arizona, the Brazilian Participation
Group, Brookhaven National Laboratory, Carnegie Mel-
lon University, University of Florida, the French Partici-
pation Group, the German Participation Group, Harvard
University, the Instituto de Astrofisica de Canarias, the
Michigan State/Notre Dame/JINA Participation Group,
Johns Hopkins University, Lawrence Berkeley National
Laboratory, Max Planck Institute for Astrophysics, Max
Planck Institute for Extraterrestrial Physics, New Mex-
ico State University, New York University, Ohio State
University, Pennsylvania State University, University of
Portsmouth, Princeton University, the Spanish Partici-
pation Group, University of Tokyo, University of Utah,
Vanderbilt University, University of Virginia, University
of Washington, and Yale University.
B.T.M. is supported by the National Science Foun-
dation Graduate Research Fellowship under Grant No.
DGE-ĂŘ1144469.
is supported by generous
grants from the David and Lucile Packard Foundation
and the Alfred P. Sloan Foundation.
D.F.M. and D.W.H. were partially supported by the
National Science Foundation (grant IIS-1124794), the
National Aeronautics and Space Administration (grant
NNX12AI50G), and the Moore-Sloan Data Science En-
vironment at NYU.
T.D.M. is supported by the National Aeronautics and
J.A.J.
Space Administration (grant NNX14AE11G).
Facilities: Kepler, PO:Hale (PHARO), FLWO:1.5m,
IRTF:SpeX
REFERENCES
Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2012, ApJS,
Burke, C. J., Bryson, S. T., Mullally, F., et al. 2014, ApJS, 210,
Aldering, G., Adam, G., Antilogus, P., et al. 2002, in Proc. SPIE,
Campante, T. L., Barclay, T., Swift, J. J., et al. 2015, ApJ, 799,
203, 21
ed. J. A. Tyson & S. Wolff, Vol. 4836, 61–72
Aldering, G., Antilogus, P., Bailey, S., et al. 2006, ApJ, 650, 510
Allard, F., Homeier, D., Freytag, B., et al. 2013, Memorie della
Societa Astronomica Italiana Supplementi, 24, 128
Armstrong, D. J., Veras, D., Barros, S. C. C., et al. 2015, ArXiv
e-prints, arXiv:1503.00692
Bacon, R., Copin, Y., Monnet, G., et al. 2001, MNRAS, 326, 23
Barentsen, G. 2015, K2flix: Kepler pixel data visualizer,
Astrophysics Source Code Library, ascl:1503.001
Bastien, F. A., Stassun, K. G., & Pepper, J. 2014, ApJ, 788, L9
Batalha, N. M., Rowe, J. F., Bryson, S. T., et al. 2013, ApJS,
204, 24
Boisse, I., Bouchy, F., Hébrard, G., et al. 2011, A&A, 528, A4
Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977
Borucki, W. J., Koch, D. G., Basri, G., et al. 2011a, ApJ, 728, 117
—. 2011b, ApJ, 736, 19
Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Tamura, M. 2015,
19
170
891
362
147, 119
ApJ, 804, 10
Charbonneau, D., Berta, Z. K., Irwin, J., et al. 2009, Nature, 462,
Coughlin, J. L., Thompson, S. E., Bryson, S. T., et al. 2014, AJ,
Crossfield, I. J. M., Petigura, E., Schlieder, J. E., et al. 2015,
Cushing, M. C., Vacca, W. D., & Rayner, J. T. 2004, PASP, 116,
Cutri, R. M., Wright, E. L., Conrow, T., & et al. 2013, VizieR
Online Data Catalog, 2328, 0
Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, 2MASS
All Sky Catalog of point sources.
Davenport, J. R. A., Ivezić, Ž., Becker, A. C., et al. 2014,
MNRAS, 440, 3430
Dekany, R., Roberts, J., Burruss, R., et al. 2013, ApJ, 776, 130
Désert, J.-M., Charbonneau, D., Torres, G., et al. 2015, ApJ, 804,
Boyajian, T. S., von Braun, K., van Belle, G., et al. 2012, ApJ,
Díaz, R. F., Almenara, J. M., Santerne, A., et al. 2014, MNRAS,
Brown, T. M., Latham, D. W., Everett, M. E., & Esquerdo, G. A.
Dotter, A., Chaboyer, B., Jevremović, D., et al. 2008, ApJS, 178,
Buchhave, L. A., Bakos, G. Á., Hartman, J. D., et al. 2010, ApJ,
Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., et al. 2014, ApJ,
ApJS, 216, 7
757, 112
2011, AJ, 142, 112
720, 1118
59
89
441, 983
790, 146
Feroz, F., Hobson, M. P., & Bridges, M. 2009, MNRAS, 398, 1601
Ferreira, P. G., & Jaffe, A. H. 2000, MNRAS, 312, 89
Fitzpatrick, E. L. 1999, PASP, 111, 63
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J.
2013, PASP, 125, 306
Mann, A. W., Feiden, G. A., Gaidos, E., Boyajian, T., & von
Braun, K. 2015, ApJ, 804, 64
Mann, A. W., Gaidos, E., & Ansdell, M. 2013b, ApJ, 779, 188
McCaughrean, M. J., & Stauffer, J. R. 1994, AJ, 108, 1382
Montet, B. T., Johnson, J. A., Muirhead, P. S., et al. 2015, ApJ,
Foreman-Mackey, D., Montet, B. T., Hogg, D. W., et al. 2015,
800, 134
ArXiv e-prints, arXiv:1502.04715
Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, 81
Gaidos, E., Mann, A. W., Lépine, S., et al. 2014, MNRAS, 443,
Goodman, J., & Weare, J. 2010, Communications in Applied
Mathematics and Computational Science, 5, 65
Grunblatt, S. K., Howard, A. W., & Haywood, R. D. 2015, ArXiv
12
2561
105
e-prints, arXiv:1501.00369
e-prints, arXiv:1503.02110
Hayden, M. R., Bovy, J., Holtzman, J. A., et al. 2015, ArXiv
217, 31
Hayward, T. L., Brandl, B., Pirger, B., et al. 2001, PASP, 113,
ApJ, 800, 85
Haywood, R. D., Collier Cameron, A., Queloz, D., et al. 2014,
MNRAS, 443, 2517
Henden, A., & Munari, U. 2014, Contributions of the
Astronomical Observatory Skalnate Pleso, 43, 518
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS,
Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126, 398
Huber, D., Silva Aguirre, V., Matthews, J. M., et al. 2014, ApJS,
Kipping, D. M. 2010, MNRAS, 408, 1758
Lantz, B., Aldering, G., Antilogus, P., et al. 2004, in Society of
Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, Vol. 5249, Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference Series, ed. L. Mazuray,
P. J. Rogers, & R. Wartmann, 146–155
Lissauer, J. J., Ragozzine, D., Fabrycky, D. C., et al. 2011, ApJS,
201, 15
211, 2
197, 8
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Mann, A. W., Brewer, J. M., Gaidos, E., Lépine, S., & Hilton,
E. J. 2013a, AJ, 145, 52
Morton, T. D. 2012, ApJ, 761, 6
—. 2015a, isochrones: Stellar model grid package, Astrophysics
Source Code Library, ascl:1503.010
—. 2015b, VESPA: False positive probabilities calculator,
Astrophysics Source Code Library, ascl:1503.011
Morton, T. D., & Johnson, J. A. 2011, ApJ, 738, 170
Morton, T. D., & Swift, J. 2014, ApJ, 791, 10
Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015, ApJS,
Newton, E. R., Charbonneau, D., Irwin, J., & Mann, A. W. 2015,
Ochsenbein, F., Bauer, P., & Marcout, J. 2000, A&AS, 143, 23
Rayner, J. T., Toomey, D. W., Onaka, P. M., et al. 2003, PASP,
Rogers, L. A. 2015, ApJ, 801, 41
Rowe, J. F., Coughlin, J. L., Antoci, V., et al. 2015, ApJS, 217, 16
Schlafly, E. F., & Finkbeiner, D. P. 2011, ApJ, 737, 103
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500,
Swift, J. J., Johnson, J. A., Morton, T. D., et al. 2013, ApJ, 764,
Thompson, B., Frinchaboy, P., Kinemuchi, K., Sarajedini, A., &
Cohen, R. 2014, AJ, 148, 85
Torres, G., Fressin, F., Batalha, N. M., et al. 2011, ApJ, 727, 24
Vacca, W. D., Cushing, M. C., & Rayner, J. T. 2003, PASP, 115,
Vanderburg, A., Montet, B. T., Johnson, J. A., et al. 2015, ApJ,
Zacharias, N., Finch, C. T., Girard, T. M., et al. 2012, VizieR
Online Data Catalog, 1322, 0
115, 362
525
105
389
800, 59
3
3
W
7
1
6
3
2
0
2
4
0
2
1
1
1
2
0
0
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
6
1
1
6
5
8
2
0
3
6
2
1
7
5
3
3
.
.
.
.
.
.
.
.
0
1
1
1
0
1
8
9
1
1
1
1
1
1
—
5
9
2
9
2
0
1
1
.
.
.
.
0
0
0
0
±
±
±
±
0
3
5
5
6
5
9
3
.
.
.
.
1
0
0
1
1
1
1
1
—
—
—
—
—
2
1
0
1
0
5
1
1
1
1
1
0
.
.
.
.
.
.
0
0
0
0
0
0
±
±
±
±
±
±
3
6
9
0
5
2
9
8
5
8
5
8
.
.
.
.
.
.
0
0
0
0
0
9
1
1
1
1
1
2
1
.
0
±
6
8
.
0
1
0
7
6
3
6
4
1
0
4
0
0
0
.
.
.
.
.
.
0
0
0
0
0
0
±
±
±
±
±
±
7
2
4
5
8
4
6
1
3
5
9
7
.
.
.
.
.
.
0
0
2
8
9
9
1
1
1
3
2
W
2
2
2
3
2
2
3
2
2
2
2
2
4
3
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
6
0
0
4
1
5
8
0
3
5
9
7
8
6
6
3
1
1
1
7
5
9
9
5
7
7
7
7
2
7
7
2
1
8
5
9
5
1
7
6
3
6
8
3
4
6
1
2
2
7
5
0
4
6
5
6
8
7
8
4
3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
2
2
2
0
2
1
0
0
1
0
3
2
1
0
1
1
2
0
1
1
1
0
1
8
9
9
9
8
9
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
3
1
W
2
2
2
2
2
2
3
2
2
2
2
2
3
2
2
2
2
2
2
3
2
2
3
2
2
2
2
2
2
2
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
3
2
6
6
1
8
2
4
3
8
9
4
4
4
8
1
0
5
6
7
2
0
9
3
1
7
2
4
2
4
8
7
2
0
4
8
8
5
1
7
5
2
6
9
3
3
6
1
3
1
6
5
0
4
6
5
7
8
7
8
4
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
2
2
2
0
2
1
0
0
1
0
3
2
1
0
1
1
2
0
1
1
1
0
1
8
9
9
9
8
9
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
K
2
2
2
3
3
2
3
3
2
2
2
2
7
3
2
2
3
2
3
2
2
2
2
2
2
2
3
2
2
2
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
0
1
9
9
7
7
1
7
0
1
5
4
8
3
2
0
6
9
4
8
6
7
0
9
7
0
3
0
8
6
7
8
3
4
0
7
0
6
1
9
6
3
6
0
4
4
7
1
4
2
7
5
0
6
6
5
9
9
8
8
5
3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
2
2
2
1
2
1
0
0
1
0
4
2
1
0
1
1
2
0
1
1
1
0
1
9
8
9
9
8
9
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
H
2
2
4
3
2
3
2
2
2
4
2
2
2
5
3
2
3
2
2
3
2
2
2
2
2
2
3
2
3
4
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
4
3
1
5
8
7
7
6
9
1
6
5
1
3
6
9
1
1
2
6
3
9
6
0
5
3
7
7
0
8
9
4
2
2
5
8
0
7
2
0
7
4
7
1
5
5
7
2
7
3
7
6
0
7
7
7
1
8
9
8
4
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
2
2
2
1
2
1
1
0
1
0
4
2
1
0
1
1
2
0
1
1
1
0
1
9
9
9
8
9
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
J
3
2
3
2
2
3
3
2
3
3
2
2
2
2
2
4
2
2
2
2
2
3
3
2
2
3
3
2
3
3
2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
8
9
3
6
7
7
7
2
7
2
8
7
6
9
8
0
4
6
0
5
3
5
5
4
5
2
7
6
2
9
4
0
4
1
0
7
2
5
4
5
7
9
8
0
3
7
2
1
1
6
4
8
0
9
4
4
0
3
7
4
9
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
1
0
0
3
2
3
1
3
1
1
0
1
1
0
4
3
2
1
1
2
2
1
1
1
2
1
2
9
9
9
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
i
9
8
1
4
1
2
5
3
6
1
6
9
8
0
0
8
1
8
5
5
2
5
7
4
5
7
7
5
1
2
2
0
0
0
0
0
0
0
0
0
0
0
0
0
1
2
0
0
0
0
0
0
1
0
0
0
0
1
0
2
0
1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
1
6
5
2
5
4
3
4
6
7
3
2
6
8
3
8
4
5
5
7
7
5
3
6
5
4
8
9
1
9
9
6
6
4
3
9
1
9
2
8
8
5
9
4
8
1
4
3
4
5
1
5
3
0
8
8
3
9
7
0
1
8
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
2
1
2
1
0
4
3
4
2
4
2
1
3
2
2
1
6
3
4
2
2
4
4
2
1
2
0
3
2
1
3
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
r
7
4
3
4
1
4
1
6
2
1
1
3
8
2
3
8
1
3
3
2
1
3
2
4
5
6
2
2
9
1
7
0
0
0
0
0
0
0
0
0
0
0
0
0
1
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
7
6
6
8
2
5
8
0
8
2
3
6
5
4
1
6
1
9
5
7
6
8
9
4
2
8
8
3
8
9
9
9
8
9
5
1
3
2
4
1
6
7
0
2
1
2
7
9
8
0
3
7
1
2
0
9
4
5
3
0
4
2
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
2
2
2
1
1
4
4
4
3
5
2
2
4
3
2
1
6
3
5
2
2
5
4
3
1
2
1
4
2
1
4
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
g
5
5
6
5
7
4
4
4
1
1
3
3
6
7
4
3
1
7
0
2
2
2
4
6
1
5
3
5
3
4
4
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
2
0
3
3
3
1
9
4
2
0
2
6
8
0
4
5
9
5
3
6
3
9
8
4
1
7
2
1
4
6
7
2
2
1
5
0
1
8
0
0
2
6
5
7
7
4
2
1
9
0
3
7
1
1
5
9
8
6
4
2
5
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3
4
4
2
1
5
5
4
4
7
3
5
2
3
2
2
6
8
4
3
3
5
2
6
3
2
2
5
2
2
5
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
t
s
e
r
e
t
n
I
f
o
1
E
L
B
A
T
s
t
c
e
j
b
O
l
l
a
r
o
f
y
r
t
e
m
o
t
o
h
P
3
2
3
4
1
3
2
1
3
0
0
0
0
0
0
1
0
0
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
1
3
1
5
7
1
9
7
1
6
0
2
6
1
9
1
7
9
.
.
.
.
.
.
.
.
.
4
2
3
2
2
4
2
1
4
1
1
1
1
1
1
1
1
1
—
1
5
2
1
0
0
0
0
.
.
.
.
0
0
0
0
±
±
±
±
3
7
4
0
4
3
6
0
.
.
.
.
5
4
2
3
1
1
1
1
—
3
5
4
1
1
1
5
4
1
1
4
5
1
9
2
7
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
1
0
2
6
7
5
6
7
3
1
0
6
6
9
2
5
3
5
5
7
2
6
6
5
5
3
9
8
2
3
4
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3
3
3
1
1
4
4
4
3
6
2
4
2
3
2
2
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
2
4
5
6
7
4
3
5
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
3
8
0
0
2
0
8
2
3
7
4
9
3
5
3
7
8
2
.
.
.
.
.
.
.
.
.
5
2
3
3
3
6
2
2
6
1
1
1
1
1
1
1
1
1
—
1
6
7
2
0
0
0
0
.
.
.
.
0
0
0
0
±
±
±
±
8
6
1
0
4
5
5
8
.
.
.
.
6
5
3
3
1
1
1
1
—
4
6
6
4
1
6
4
5
3
6
3
2
9
1
1
8
0
0
0
0
0
0
0
0
0
1
0
0
0
0
1
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
2
1
2
8
1
5
9
7
7
4
1
4
9
1
4
0
3
0
8
4
8
6
4
2
5
7
6
3
9
2
1
9
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
4
5
4
2
1
5
5
5
4
7
3
6
2
4
3
2
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
3
4
8
0
2
1
0
2
1
6
4
7
5
2
1
0
2
2
1
3
5
9
2
1
0
2
8
0
5
8
3
3
1
0
2
5
6
0
7
6
3
1
0
2
2
3
2
4
8
3
1
0
2
8
9
0
3
9
3
1
0
2
6
4
4
3
0
4
1
0
2
2
9
3
5
4
4
1
0
2
1
0
5
5
6
4
1
0
2
0
5
3
5
0
5
1
0
2
3
8
2
6
4
5
1
0
2
0
6
8
9
4
5
1
0
2
3
8
8
5
5
5
1
0
2
3
1
0
5
6
5
1
0
2
3
8
4
9
6
5
1
0
2
5
3
0
7
7
5
1
0
2
6
1
3
6
9
5
1
0
2
3
2
0
3
1
6
1
0
2
5
8
9
7
1
6
1
0
2
0
5
6
9
2
6
1
0
2
9
6
5
5
3
6
1
0
2
6
2
4
9
4
6
1
0
2
7
7
4
2
0
7
1
0
2
7
4
2
6
3
7
1
0
2
5
0
3
4
5
7
1
0
2
7
6
0
9
7
7
1
0
2
9
4
7
8
2
8
1
0
2
1
7
3
5
5
8
1
0
2
2
5
5
2
1
9
1
0
2
4
9
2
9
2
9
1
0
2
1
V
1
B
C
I
P
E
13
.
)
2
1
0
2
.
l
a
t
e
s
a
i
r
a
h
c
a
Z
(
e
u
g
o
l
a
t
a
C
4
C
A
C
U
e
h
t
n
i
d
e
t
r
o
p
e
r
s
a
)
4
1
0
2
i
r
a
n
u
M
&
n
e
d
n
e
H
(
6
R
D
)
S
S
A
P
A
(
y
e
v
r
u
S
y
k
S
-
l
l
A
c
i
r
t
e
m
o
t
o
h
P
O
S
V
A
A
e
h
t
.
)
3
0
0
2
.
l
a
t
e
i
r
t
u
C
(
s
e
c
r
u
o
S
t
n
i
o
P
f
o
g
o
l
a
t
a
C
y
k
S
-
l
l
A
S
S
A
M
2
e
h
t
.
)
3
1
0
2
.
l
a
t
e
i
r
t
u
C
(
e
s
a
e
l
e
R
a
t
a
D
e
s
i
W
L
L
A
e
h
t
m
o
r
f
m
o
r
f
m
o
r
f
e
d
u
t
i
n
g
a
M
1
e
d
u
t
i
n
g
a
M
2
e
d
u
t
i
n
g
a
M
3
.
t
e
t
n
o
m
_
s
t
e
n
a
l
p
_
2
k
/
m
o
c
.
h
p
a
r
g
r
e
t
l
fi
/
/
:
s
p
t
t
h
t
a
m
r
o
f
e
v
i
t
c
a
r
e
t
n
i
n
i
e
l
b
a
l
i
a
v
a
e
r
a
a
t
a
d
e
s
e
h
T
—
.
e
t
o
N
14
Stellar Properties for all Objects of Interest
TABLE 2
(Degrees)
—
—
—
—
—
Name
K2-4
—
—
K2-5
K2-3
K2-6
K2-7
—
K2-8
K2-9
K2-19
Alternate RA (J2000) Dec (J2000)
(Degrees)
−3.905585
−3.094936
−2.520881
−1.877976
−1.454787
−1.198477
−1.065755
−0.907261
−0.284375
0.005301
0.603575
1.230738
1.285956
1.375947
1.510249
1.577513
1.690636
1.986840
2.244884
2.321476
2.502696
2.594245
2.807619
3.681584
4.254747
4.557340
4.988131
5.894323
6.412261
7.588391
7.959611
174.745639
178.161110
174.011629
169.303502
172.334949
178.192260
167.093771
174.266345
169.793666
176.264467
174.960319
171.515164
170.103081
176.075940
176.992193
167.171300
172.121957
169.042002
173.192036
179.491659
170.155529
178.057026
177.234262
175.240794
178.110796
175.097258
168.542699
175.654343
178.329776
172.560461
174.656968
K2-10
K2-11
K2-12
—
K2-13
K2-14
K2-17
K2-18
—
K2-15
K2-16
—
—
—
—
EPIC
201208431
201257461
201295312
201338508
201367065
201384232
201393098
201403446
201445392
201465501
201505350
201546283
201549860
201555883
201565013
201569483
201577035
201596316
201613023
201617985
201629650
201635569
201649426
201702477
201736247
201754305
201779067
201828749
201855371
2019125521
201929294
Mass
(M(cid:12))
0.63+0.03−0.03
1.50+0.04−0.02
1.07+0.07−0.07
0.53+0.01−0.01
0.53+0.02−0.02
0.97+0.07−0.07
0.97+0.06−0.06
1.01+0.08−0.06
0.79+0.03−0.04
0.24+0.05−0.03
0.84+0.04−0.04
0.89+1.15−0.07
0.73+0.03−0.03
0.54+0.07−0.01
0.51+0.13−0.03
0.83+0.05−0.05
0.94+0.04−0.06
1.35+0.04−0.56
1.01+0.05−0.06
0.52+0.03−0.03
0.80+0.04−0.04
0.47+0.01−0.01
1.29+0.02−0.02
0.87+0.06−0.06
0.72+0.06−0.03
0.67+0.04−0.03
0.91+0.03−0.04
0.74+1.06−0.04
0.71+0.02−0.05
0.413+0.043
−0.043
0.73+0.06−0.09
Radius
(R(cid:12))
0.60+0.02−0.02
10.96+0.82−0.93
1.09+0.20−0.11
0.52+0.01−0.01
0.52+0.02−0.02
0.96+0.14−0.09
0.96+0.17−0.08
1.12+0.26−0.14
0.74+0.02−0.03
0.25+0.04−0.03
0.81+0.09−0.05
0.88+7.37−0.10
0.69+0.02−0.02
0.52+0.08−0.01
0.50+0.12−0.03
0.79+0.06−0.05
0.93+0.16−0.07
5.15+0.20−4.39
1.01+0.27−0.09
0.49+0.03−0.03
0.78+0.09−0.05
0.45+0.01−0.01
8.15+0.32−0.23
0.85+0.11−0.08
0.68+0.06−0.03
0.64+0.03−0.03
0.92+0.20−0.07
0.71+9.64−0.06
0.66+0.02−0.03
0.394+0.038
−0.038
0.70+0.04−0.08
Teff
(K)
[Fe/H]
(dex)
4197+45−43 −0.12+0.10−0.12
5141+38−42 −0.21+0.01−0.01
−0.02+0.15−0.18
5989+100−81
4102+45−41 −0.51+0.04−0.06
3951+33−38 −0.30+0.07−0.06
5850+79−98 −0.14+0.17−0.20
5772+72−91 −0.07+0.16−0.16
6445+81−111 −0.50+0.15−0.13
4890+38−58 −0.01+0.11−0.13
3468+20−19 −0.46+0.12−0.10
5519+49−82 −0.27+0.10−0.10
−0.09+0.31−0.15
5422+194−93
0.05+0.15−0.14
4523+43−47
4419+29−33 −0.98+0.62−0.11
−0.44+0.47−0.08
3987+142−68
5192+55−70 −0.09+0.17−0.15
5647+60−89 −0.04+0.14−0.17
5433+49−144 −0.12+0.01−0.17
5800+53−90
0.03+0.13−0.17
3742+31−36 −0.08+0.10−0.11
5698+45−82 −0.54+0.12−0.14
3789+17−16 −0.37+0.03−0.04
5086+24−26 −0.17+0.01−0.01
5618+86−85 −0.26+0.17−0.18
5131+69−65 −0.46+0.20−0.14
4761+50−57 −0.40+0.12−0.17
6166+30−51 −0.54+0.07−0.12
5552+87−97 −0.69+0.34−0.23
4320+56−47
0.15+0.09−0.22
3503+60−60
0.09+0.09−0.09
4786+48−53 −0.16+0.22−0.34
Distance
(pc)
218+11−10
1651+121−134
331+61−35
181+7−7
42+2−2
343+52−33
433+75−38
362+86−48
405+14−16
66+11−7
291+33−20
251+2138−29
249+9−9
289+46−9
506+154−38
152+12−10
271+48−21
2019+71−1728
294+78−27
111+8−9
290+34−18
219+8−8
2537+92−68
673+87−63
437+43−22
324+16−16
188+39−15
146+1996−12
134+5−6
34+4−4
197+13−24
Note. — These values and their uncertainties are derived from MULTINEST analysis and the numbers are computed as the 0.158,
0.500, and 0.842 posterior sample quantiles. The coordinates are retrieved directly from the EPIC. These data are available in
interactive form at https://filtergraph.com/k2_planets_montet.
1 Parameters inferred from spectroscopic observations.
n
o
i
t
i
s
o
p
s
i
D
)
K
(
q
e
T
)
U
A
(
a
(cid:63)
R
/
a
t
s
e
r
e
t
n
I
f
o
t
e
n
a
l
P
P
F
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
e
t
a
d
i
d
n
a
C
e
t
a
d
i
d
n
a
C
t
e
n
a
l
P
e
t
a
d
i
d
n
a
C
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
e
t
a
d
i
d
n
a
C
e
t
a
d
i
d
n
a
C
2
P
F
e
t
a
d
i
d
n
a
C
P
F
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
P
F
t
e
n
a
l
P
t
e
n
a
l
P
t
e
n
a
l
P
P
F
t
e
n
a
l
P
t
e
n
a
l
P
2
P
F
e
t
a
d
i
d
n
a
C
e
t
a
d
i
d
n
a
C
e
t
a
d
i
d
n
a
C
4
9
5
2
4
2
8
7
4
4
8
5
2
1
1
3
0
9
9
5
7
4
4
1
1
1
1
8
1
1
6
1
2
1
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
1
4
4
6
1
3
7
6
4
7
4
1
9
1
3
1
0
7
1
5
6
1
9
1
6
8
3
7
9
5
8
6
5
5
3
9
1
4
2
7
9
7
2
6
8
6
6
8
5
1
9
6
1
—
1
9
3
6
6
6
2
6
4
2
0
7
1
5
2
6
3
8
2
5
3
1
1
2
3
1
4
5
5
3
1
1
2
1
2
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
3
7
0
3
6
9
7
1
3
0
6
8
8
3
4
1
0
2
7
2
1
0
8
4
3
3
0
1
4
1
3
1
0
5
6
5
7
7
4
4
5
9
7
5
6
7
5
1
5
1
±
2
7
2
—
7
3
9
8
0
9
7
4
8
9
2
6
6
3
0
0
2
2
1
1
0
5
0
0
0
1
1
0
4
4
5
1
3
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
8
5
5
8
1
5
6
8
4
8
9
3
3
0
9
5
7
4
6
3
6
3
5
5
9
1
0
3
8
0
4
4
7
4
8
9
7
6
5
5
8
8
8
4
6
7
5
0
7
7
3
0
0
0
0
0
0
0
1
1
1
0
0
0
3
0
0
1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
—
3
8
1
1
2
0
0
7
1
5
5
3
6
1
2
1
9
2
1
2
1
5
2
9
0
4
2
1
3
6
1
2
0
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
5
8
2
0
0
5
7
0
7
7
4
6
2
4
9
9
0
1
2
6
1
7
1
9
5
2
0
8
1
6
8
7
2
9
2
6
7
8
5
1
2
6
8
5
1
6
5
3
2
0
1
0
1
1
1
1
2
0
0
0
2
0
0
1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
5
5
0
0
.
0
±
1
9
4
1
.
0
—
6
9
4
6
8
6
1
6
9
5
9
6
7
9
7
2
5
2
8
2
4
4
6
4
4
7
7
7
8
3
1
0
7
.
5
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
4
.
1
2
9
6
0
0
0
5
8
0
0
4
0
1
0
0
2
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
6
2
5
9
6
9
9
2
4
5
5
9
7
4
9
7
9
5
4
0
3
9
7
4
7
8
9
0
2
2
9
7
2
1
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
7
4
8
0
7
4
0
5
5
4
7
0
2
2
7
0
6
1
2
1
2
1
7
3
1
5
2
2
4
3
1
2
5
—
4
8
3
7
8
4
3
2
5
1
5
8
1
9
6
5
6
5
5
3
7
7
3
9
8
6
1
6
9
6
1
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
7
3
3
8
3
0
1
1
7
0
0
1
5
1
2
5
6
1
6
5
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
5
7
3
8
4
8
4
8
7
6
4
7
9
6
8
9
4
4
2
3
8
9
7
0
6
1
4
0
0
2
0
6
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
8
8
5
2
9
8
1
6
2
0
7
6
7
5
9
9
3
2
1
2
3
2
4
5
3
3
1
2
6
4
7
5
3
0
.
9
±
3
8
.
3
8
—
3
2
0
9
1
3
4
2
0
3
0
8
1
6
6
0
7
0
.
4
8
4
2
2
4
3
1
1
5
8
4
5
2
5
4
9
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
9
0
0
0
3
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
2
1
7
1
0
0
1
6
0
7
4
7
2
2
8
6
7
5
1
3
7
2
6
3
5
5
0
9
6
9
9
9
1
3
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
9
2
7
4
5
1
1
2
2
2
1
1
2
2
1
2
2
0
2
—
5
1
9
6
7
5
0
0
1
0
9
3
2
8
3
5
2
0
5
1
3
2
2
4
3
1
6
3
4
9
4
5
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
5
9
3
9
0
0
0
0
3
1
0
0
9
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
3
9
1
9
3
3
3
4
8
8
2
5
9
8
3
1
7
9
8
7
8
2
1
1
4
2
9
5
3
7
8
8
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
2
5
7
2
3
2
2
2
7
2
1
1
4
3
7
3
3
1
2
3
2
.
0
±
4
2
.
2
—
3
E
L
B
A
T
s
t
c
e
j
b
O
l
l
a
r
o
f
s
e
i
t
r
e
p
o
r
P
t
e
n
a
l
P
)
⊕
R
(
s
u
i
d
a
R
)
8
0
8
6
5
4
2
-
D
J
B
(
h
c
o
p
E
)
s
y
a
d
(
d
o
i
r
e
P
e
t
a
d
i
d
n
a
C
0
0
9
7
7
5
8
2
5
0
1
2
3
3
0
2
0
3
9
4
9
4
0
2
1
1
7
3
5
5
6
8
8
8
0
0
1
3
0
0
0
0
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
3
4
5
5
1
0
6
4
3
9
0
7
7
9
7
2
2
2
1
3
5
8
3
4
6
6
1
4
7
4
5
8
1
1
7
0
7
1
1
8
4
7
6
1
6
1
9
7
1
4
2
6
5
3
6
1
3
8
2
0
6
8
4
5
2
7
3
5
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
4
9
0
6
4
5
4
5
0
3
9
7
5
5
6
4
7
1
1
2
1
—
3
3
6
2
4
4
5
9
6
5
6
4
4
7
6
3
0
3
1
5
5
3
3
1
7
2
0
1
5
1
4
4
0
0
0
1
0
0
0
1
0
0
0
0
0
0
1
0
0
0
0
1
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
1
2
8
0
2
4
4
2
9
5
3
4
5
4
0
6
0
8
6
9
0
5
0
6
0
5
1
8
3
3
6
5
6
4
7
2
8
8
5
4
5
4
5
2
1
7
3
2
2
3
5
8
6
4
1
9
8
5
4
4
3
3
6
5
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
2
3
1
1
3
3
5
7
4
1
5
9
3
3
4
3
1
1
1
2
7
2
0
0
.
0
±
9
4
8
1
.
8
2
—
7
7
7
3
3
5
2
7
5
7
9
5
7
2
3
1
9
3
3
3
0
3
0
5
6
8
4
5
5
1
0
6
6
7
0
1
1
6
0
2
1
4
7
9
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
1
3
8
5
6
4
1
2
6
2
9
1
3
0
8
1
6
9
8
4
4
9
4
7
9
0
6
2
4
3
4
6
9
0
6
8
4
7
1
3
1
9
4
7
3
9
1
8
4
4
7
0
4
5
4
4
5
5
7
3
7
0
1
7
0
6
3
5
9
4
0
6
9
1
3
6
9
2
0
9
7
6
0
7
6
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
0
8
0
4
0
8
0
0
9
0
7
5
6
5
5
5
1
1
1
1
2
3
2
1
1
1
5
—
0
2
2
2
4
6
8
7
7
9
5
9
4
0
0
8
5
3
1
9
0
6
0
7
2
2
9
1
0
2
6
7
1
2
0
4
2
2
0
4
1
2
0
0
0
0
0
0
0
0
0
0
0
0
0
2
0
3
0
0
0
0
0
0
0
0
0
0
0
0
0
3
0
2
0
0
0
0
0
0
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
±
9
3
3
6
0
0
1
5
7
8
7
7
2
1
2
0
7
6
5
3
4
2
9
4
6
8
6
8
1
6
0
1
2
5
7
5
0
6
6
0
7
4
0
6
2
1
8
8
4
1
6
7
1
3
0
7
3
1
2
9
8
8
6
3
2
5
9
0
8
7
3
7
9
9
6
7
2
2
3
6
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
9
7
1
9
9
9
3
7
0
7
7
8
8
7
8
5
3
2
1
1
1
1
3
3
4
2
1
8
2
0
0
.
0
±
8
8
4
4
9
.
2
3
—
b
4
-
2
K
/
1
0
.
1
3
4
8
0
2
1
0
2
1
0
.
1
6
4
7
5
2
1
0
2
c
5
-
2
K
/
1
0
.
8
0
5
8
3
3
1
0
2
b
5
-
2
K
/
2
0
.
8
0
5
8
3
3
1
0
2
b
3
-
2
K
/
1
0
.
5
6
0
7
6
3
1
0
2
c
3
-
2
K
/
2
0
.
5
6
0
7
6
3
1
0
2
b
6
-
2
K
/
1
0
.
2
3
2
4
8
3
1
0
2
b
7
-
2
K
/
1
0
.
8
9
0
3
9
3
1
0
2
1
0
.
2
1
3
5
9
2
1
0
2
b
8
-
2
K
/
1
0
.
2
9
3
5
4
4
1
0
2
1
0
.
6
4
4
3
0
4
1
0
2
c
9
1
-
2
K
/
1
0
.
0
5
3
5
0
5
1
0
2
b
9
1
-
2
K
/
2
0
.
0
5
3
5
0
5
1
0
2
b
9
-
2
K
/
1
0
.
1
0
5
5
6
4
1
0
2
2
0
.
2
9
3
5
4
4
1
0
2
b
0
1
-
2
K
/
1
0
.
5
3
0
7
7
5
1
0
2
b
1
1
-
2
K
/
1
0
.
6
1
3
6
9
5
1
0
2
b
2
1
-
2
K
/
1
0
.
3
2
0
3
1
6
1
0
2
1
0
.
3
8
4
9
6
5
1
0
2
b
3
1
-
2
K
/
1
0
.
0
5
6
9
2
6
1
0
2
b
4
1
-
2
K
/
1
0
.
9
6
5
5
3
6
1
0
2
1
0
.
5
8
9
7
1
6
1
0
2
1
0
.
6
2
4
9
4
6
1
0
2
1
0
.
3
8
2
6
4
5
1
0
2
1
0
.
0
6
8
9
4
5
1
0
2
1
0
.
3
8
8
5
5
5
1
0
2
1
0
.
3
1
0
5
6
5
1
0
2
b
5
1
-
2
K
/
1
0
.
7
4
2
6
3
7
1
0
2
c
6
1
-
2
K
/
1
0
.
5
0
3
4
5
7
1
0
2
b
6
1
-
2
K
/
2
0
.
5
0
3
4
5
7
1
0
2
1
0
.
7
7
4
2
0
7
1
0
2
1
0
.
7
6
0
9
7
7
1
0
2
1
b
8
1
-
2
K
/
1
0
.
2
5
5
2
1
9
1
0
2
b
7
1
-
2
K
/
1
0
.
1
7
3
5
5
8
1
0
2
1
0
.
9
4
7
8
2
8
1
0
2
1
0
.
4
9
2
9
2
9
1
0
2
15
)
2
.
5
§
e
e
s
(
s
c
i
t
a
m
e
t
s
y
s
g
n
i
l
e
d
o
m
e
s
i
o
n
o
t
e
u
d
e
v
i
t
i
s
o
p
e
s
l
a
f
a
d
e
r
a
l
c
e
D
.
s
n
o
i
t
a
v
r
e
s
b
o
c
i
p
o
c
s
o
r
t
c
e
p
s
m
o
r
f
d
e
r
r
e
f
n
i
s
r
e
t
e
m
a
r
a
P
1
2
e
v
i
t
c
a
r
e
t
n
i
n
i
e
l
b
a
l
i
a
v
a
e
r
a
a
t
a
d
e
s
e
h
T
.
s
g
n
i
l
p
m
a
s
r
o
i
r
e
t
s
o
p
C
M
C
M
f
o
n
o
i
t
a
i
v
e
d
d
r
a
d
n
a
t
s
d
n
a
n
a
e
m
e
h
t
y
b
n
e
v
i
g
e
r
a
s
e
i
t
n
i
a
t
r
e
c
n
u
d
n
a
s
e
u
l
a
v
e
s
e
h
T
—
.
e
t
o
N
.
t
e
t
n
o
m
_
s
t
e
n
a
l
p
_
2
k
/
m
o
c
.
h
p
a
r
g
r
e
t
l
fi
/
/
:
s
p
t
t
h
t
a
m
r
o
f
16
Detected Companions to Candidate Host Stars
TABLE 4
Primary
Aperture1
(arcsec)
RA2
(J2000)
Dec2
(J2000)
Detection3
Separation4
(arcsec)
∆r5
(mag)
Max Depth6 Observed Depth7
(ppt)
(ppt)
SDSS
SDSS
SDSS
2.3
62.3
SDSS
SDSS
SDSS/AO
SDSS
171.515265
170.097556
1.229950
1.288007
SDSS/AO
SDSS
5.6
4.8
0.8
9.1
2.1
7.5
172.118116
1.687798
178.195303
-1.192501
174.267663
-0.909645
5.87 ± 0.06
2.26 ± 0.03
2.98 ± 0.05a
21.21 ± 0.05b
-3.902146
-3.093431
-2.522528
-1.873647
174.748988
178.164376
174.010158
169.308176
5.90 ± 0.12
5.04 ± 0.03
7.10 ± 0.10
4.35 ± 0.03
5.93 ± 0.03
4.56 ± 0.08
17.25 ± 0.15b
12.91 ± 0.18b
8.12 ± 0.09b
22.92 ± 0.07b
24.14 ± 0.06b
9.78 ± 0.14b
15.9
19.9
11.9
15.9
19.9
13.9
15.9
15.9
13.9
11.9
19.9
17.9
13.9
10.0
10.0
19.9
19.9
15.9
19.9
15.9
15.9
11.9
19.9
10.0
13.9
11.9
19.9
11.9
19.9
13.9
19.9
201208431
201257461
201295312
201338508
201367065
201384232
201393098
201403446
201445392
201465501
201505350
201546283
201549860
201555883
201565013
201569483
201577035
201596316
201613023
201617985
201629650
201635569
201649426
201702477
201736247
201754305
201779067
201828749
201855371
201912552
201929294
1 Defined aperture used to create the K2 stellar light curve.
2 Position of imaged companion.
3 Dataset used to detect the imaged companion.
4 Distance between the primary K2 target star and companion, in the dataset in which the companion is detected.
5 Difference in r-band magnitude between the primary K2 target star and the companion.
6 Observed “transit” depth if the imaged companion’s flux were fully contained in the aperture and if it were an equal-mass eclipsing binary, leading
to an eclipse depth of 50%. This is the maximum possible false positive eclipse depth, as described in §4.3.
7 Observed transit depth in the K2 dataset. If larger than the "max depth," this transit event cannot be caused by eclipses of the background star.
a Separation from AO imaging
b Separation from SDSS photometry
c ∆r inferred from JHK relative photometry.
1.20
30.54
0.30
1.07
1.26
0.68
0.53
0.23
0.78
2.83
2.64
2.33
0.80
3.50
45.8
160
1.44
0.70
0.42
1.10
0.58
9.43
216
6.70
1.21
0.80
84.9
0.76
0.99
2.85
13.56
12.15 ± 0.12b
4.65 ± 0.09
12.30 ± 0.14b
5.98 ± 0.06
17.19 ± 0.12b
5.40 ± 0.03
175.238916
3.678764
170.158905
2.502107
175.645724
5.894714
2.46 ± 0.04a
2.0 ± 0.1c
3.5
2.0
6.9
SDSS
SDSS
AO
137
False Positive Probability Calculation Results
TABLE 5
PrBEB
17
3
fp
FPP
Disposition
1.000
0.019
–
0.026
–
–
Planet
FP
Candidate
δsec,max [ppt]1 AO?2
Planet
Candidate
Planet
Planet
Planetc
Planetc
Planet
Planet
-
-
Y
-
-
-
-
-
-
Y
-
-
-
-
-
-
-
-
-
-
Y
-
Y
-
-
-
-
-
-
-
-
-
Y
-
Y
-
PrHEB
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
PrEB
< 10−4
0.998
1.4 × 10−4
< 10−4
< 10−4
< 10−4
< 10−4
8.4 × 10−3
< 10−4
4.8 × 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
7.0 × 10−4
< 10−4
0.51
0.59
0.04
0.63
0.33
0.15
0.67
0.44
0.52
0.18
0.26
0.18
0.68
2.69
0.70
0.15
0.18
0.94
1.69
2.06
0.14
0.45
0.08
0.27
0.43
0.79
3.10
0.70
0.42
0.65
0.38
1.97
0.39
0.62
0.47
3.12
8.1 × 10−4
1.7 × 10−3
< 10−4
2.9 × 10−3
1.7 × 10−4
1.1 × 10−4
< 10−4
< 10−4
1.1 × 10−3
< 10−4
2.1 × 10−3
5.8 × 10−3
5.6 × 10−3
1.6 × 10−4
2.6 × 10−4
0.019
8.1 × 10−4
1.4 × 10−4 Candidatea
2.9 × 10−3
1.7 × 10−4
1.1 × 10−4
< 10−4
8.5 × 10−3
1.1 × 10−3
4.9 × 10−4 Candidatea
2.1 × 10−3
5.8 × 10−3
Planet
5.6 × 10−3
Planetd
1.7 × 10−4
Planetd
9.6 × 10−4 Candidatea
Candidate
201208431.01/K2-4b
201257461.01
201295312.01
201338508.01/K2-5c
201338508.02/K2-5b
201367065.01/K2-3b
201367065.02/K2-3c
201384232.01/K2-6b
201393098.01/K2-7b
201403446.01
201445392.01/K2-8b
201445392.02
201465501.01/K2-9b
201505350.01/K2-19c
201505350.02/K2-19b
201546283.01
201549860.01
201555883.01
201565013.01
201569483.01
201577035.01/K2-10b
201596316.01/K2-11b
201613023.01/K2-12b
201617985.01
201629650.01/K2-13b
201635569.01/K2-14b
201649426.01
201702477.01
201736247.01/K2-15b
201754305.01/K2-16c
201754305.02/K2-16b
201779067.01
201828749.01
201855371.01/K2-17b
201912552.01/K2-18b
201929294.01
Note. — Results of the vespa astrophysical false positive probability calculations for all candidates. Likely false positives (FPP
> 0.9, or otherwise designated) are marked in red. Candidates are declared to be validated planets if FPP < 0.01. EB, BEB,
and HEB refer to the three considered astrophysical false positive scenarios, and the relative probability of each is listed in the
appropriate column. Planets previously identified in the literature are marked.
1 Maximum depth of potential secondary eclipse signal.
2 Whether adaptive optics observation is presented in this paper.
3 Integrated planet occurrence rate assumed between 0.7× and 1.3× the candidate radius
a Despite low FPP, returned to candidate status out of abundance of caution due to secondary star detection within or near
photometric aperture.
b Declared a false positive due to noise modeling systematics (see §5.2).
c Identified as planets by Crossfield et al. (2015).
d Identified as planets by Armstrong et al. (2015).
7.3 × 10−3
< 10−4
< 10−4
1.2 × 10−3
< 10−4
0.012
2.0 × 10−4
4.9 × 10−3
< 10−4
1.2 × 10−3
2.1 × 10−4
1.4 × 10−3
9.9 × 10−4
1.3 × 10−3
3.8 × 10−4
8.7 × 10−3
< 10−4
0.063
0.174
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
< 10−4
0.104
6.6 × 10−3
< 10−4
< 10−4
< 10−4
7.2 × 10−3
< 10−4
< 10−4
< 10−4
0.783
0.822
4.4 × 10−4
< 10−4
< 10−4
< 10−4
5.9 × 10−4
< 10−4
0.896
0.137
4.8 × 10−4
1.0 × 10−4
2.3 × 10−4
6.9 × 10−4
1.5 × 10−3
1.2 × 10−3
0.976
0.645
8.7 × 10−3
< 10−4
0.21
0.00
0.17
0.22
0.22
0.22
0.16
0.07
0.05
0.19
0.18
0.21
0.21
0.04
0.07
0.00
0.04
–
0.07
0.00
0.07
0.06
0.18
0.18
0.13
0.05
0.00
0.05
0.19
0.21
0.19
0.00
0.01
0.01
0.21
–
4.4 × 10−4
1.2 × 10−3
< 10−4
0.012
7.8 × 10−4
4.9 × 10−3
0.968
0.644
< 10−4
< 10−4
FP
Planet
Planet
Planet
0.026
–
0.853
0.996
Planet
Planet
FP
Planet
Planet
Planet
FP
Planet
Planet
FPb
FPb
Candidate
–
–
–
1.000
0.145
Candidate
Candidate
–
Candidate
|
0909.1435 | 2 | 0909 | 2009-09-10T08:41:29 | Disc-planet interactions in sub-keplerian discs | [
"astro-ph.EP"
] | One class of protoplanetary disc models, the X-wind model, predicts strongly subkeplerian orbital gas velocities, a configuration that can be sustained by magnetic tension. We investigate disc-planet interactions in these subkeplerian discs, focusing on orbital migration for low-mass planets and gap formation for high-mass planets. We use linear calculations and nonlinear hydrodynamical simulations to measure the torque and look at gap formation. In both cases, the subkeplerian nature of the disc is treated as a fixed external constraint. We show that, depending on the degree to which the disc is subkeplerian, the torque on low-mass planets varies between the usual Type I torque and the one-sided outer Lindblad torque, which is also negative but an order of magnitude stronger. In strongly subkeplerian discs, corotation effects can be ignored, making migration fast and inward. Gap formation near the planet's orbit is more difficult in such discs, since there are no resonances close to the planet accommodating angular momentum transport. In stead, the location of the gap is shifted inwards with respect to the planet, leaving the planet on the outside of a surface density depression. Depending on the degree to which a protoplanetary disc is subkeplerian, disc-planet interactions can be very different from the usual Keplerian picture, making these discs in general more hazardous for young planets. | astro-ph.EP | astro-ph |
Astronomy&Astrophysicsmanuscript no. 13184
November 5, 2018
c(cid:13) ESO 2018
Letter to the Editor
Disc-planet interactions in subkeplerian discs
S.-J. Paardekooper
Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Centre for Mathematical Sciences,
Wilberforce Road, Cambridge CB3 0WA, United Kingdom
e-mail: [email protected]
Draft version November 5, 2018
ABSTRACT
Context. One class of protoplanetary disc models, the X-wind model, predicts strongly subkeplerian orbital gas velocities, a configu-
ration that can be sustained by magnetic tension.
Aims.We investigate disc-planet interactions in these subkeplerian discs, focusing on orbital migration for low-mass planets and gap
formation for high-mass planets.
Methods. We use linear calculations and nonlinear hydrodynamical simulations to measure the torque and look at gap formation. In
both cases, the subkeplerian nature of the disc is treated as a fixed external constraint.
Results.We show that, depending on the degree to which the disc is subkeplerian, the torque on low-mass planets varies between the
usual Type I torque and the one-sided outer Lindblad torque, which is also negative but an order of magnitude stronger. In strongly
subkeplerian discs, corotation effects can be ignored, making migration fast and inward. Gap formation near the planet's orbit is more
difficult in such discs, since there are no resonances close to the planet accommodating angular momentum transport. The location of
the gap is shifted inwards with respect to the planet, leaving the planet on the outside of a surface density depression.
Conclusions. Depending on the degree to which a protoplanetary disc is subkeplerian, disc-planet interactions can be very different
from the usual Keplerian picture, making these discs in general more hazardous for young planets.
Key words. planets and satellites: formation -- planetary systems: protoplanetary discs
1. Introduction
Planets form in circumstellar discs, and the gravitational interac-
tion between planet and disc plays a major role in shaping plan-
etary systems. Tidal waves excited by the planet lead to orbital
migration (Goldreich & Tremaine 1979), which comes in sev-
eral flavours. Type I migration (Ward 1997) is thought to hold for
low-mass planets, up to a few times the mass of the Earth (M⊕).
High-mass planets, comparable to Jupiter (with mass MJ), can
tidally truncate the disc and open up a deep annular gap around
their orbit. The resulting Type II migration is driven by the vis-
cous evolution of the disc (Lin & Papaloizou 1986). A strongly
dynamical form of migration, Type III (Masset & Papaloizou
2003), can be achieved for intermediate-mass planets that par-
tially open up a gap in massive discs, or by planets placed
on a strong density gradient (Pepli´nski et al. 2008). For a re-
cent overview of disc-planet interactions, see Papaloizou et al.
(2007).
Most studies of disc-planet interactions have focused on
discs that have a Keplerian velocity profile, possibly with a slight
correction for a radial pressure gradient. One class of protoplan-
etary disc models, however, predicts strongly subkeplerian mo-
tion due to magnetic tension (Shu et al. 2007). In an interesting
study, Adams et al. (2009) point out that in these X-wind discs,
migration due to the strong head wind experienced by the planet,
which is moving with Keplerian velocity, can overcome Type I
migration for Earth-sized planets. On the other hand, gap for-
mation should not be affected that much, with the planet located
slightly towards the inner edge of the gap.
In this Letter, we study disc-planet interactions in subkeple-
rian discs, under the simplifying assumption that the subkeple-
rian nature of the disc is an external and fixed constraint. This
allows us to use purely hydrodynamical models as a first ap-
proximation. Future models should self-consistently include the
evolution of the magnetic field. We outline the numerical and
physical set-up in Sect. 2, present the results in Sect. 3 and con-
clude in Sect. 4.
2. Disc model
The basic equations are conservation of mass, momentum, and
energy in a 3D cylindrical geometry (r, ϕ, z). We make two sim-
plifications at this point: by considering an isothermal equation
of state, we remove the need to solve the energy equation, and
by integrating the resulting system vertically, we end up with
a two-dimensional problem. Although it has been pointed out
that the isothermal approximation is not valid in the inner re-
gions of protoplanetary discs (Paardekooper & Mellema 2006a,
2008), we see below that in strongly subkeplerian discs these
effects play essentially no role. The sound speed is given by
cs = HΩ0, where H is the scale height of the disc and Ω0 the
equilibrium angular velocity (see below). We varied h = H/rp,
where rp is the orbital radius of the planet, between 0.05 and
0.2. The higher values of H are appropriate if the disc close to
the star is 'puffed up' by direct irradiation. A kinematic viscos-
ity ν was used, with ν(r) such that the initial, constant surface
density is a stationary solution with no accretion. In the text,
we quote the corresponding α-parameter at the location of the
planet, α = ν(rp)/(H2Ωp). The resulting set of equations can
be found elsewhere (see Paardekooper & Papaloizou 2009). The
value of the surface density at r = rp, Σp, can be chosen arbitrar-
2
S.-J. Paardekooper: Disc-planet interactions in subkeplerian discs
f=1
f=0.8
f=0.6
f=0.4
100
0
Γ
/
Γ
10
h=0.05
h=0.1
h=0.2
0.6
0.8
1.0
r/rp
1.2
1.4
Fig. 1. Position of the inner (open triangles) and outer (solid cir-
cles) Lindblad resonances in discs with varying f . Pressure ef-
fects have been ignored in calculating the resonance positions.
ily in non-selfgravitating discs; we normalise the torque accord-
ingly (see Sect. 3.1).
The gravitational potential entering the equations contains
terms due to the central star and the direct and indirect com-
ponent of the planet potential. We use a softened point mass
potential for the planet, with a softening parameter b = 0.6H
(see Paardekooper & Papaloizou 2009). This value of b is ap-
propriate to account for 3D effects. The planet is assumed to be
on a fixed circular orbit. For the central star potential we use
Φ∗ = − f 2GM∗/r, where f is a dimensionless parameter govern-
ing the degree to which the disc is subkeplerian. It is easy to see
that the resulting equilibrium angular velocity, neglecting pres-
sure effects, is given by Ω0 = f ΩK, where ΩK is the Keplerian
angular velocity. For f = 1, we have a purely Keplerian disc,
while for f < 1, the disc is subkeplerian. X-wind models sug-
gest f ≈ 0.6, or even f ≈ 0.34 during FU Orionis outbursts
(Shu et al. 2007).
We combine a linear code (Paardekooper & Papaloizou
2009) with fully nonlinear hydrodynamic calculations with
RODEO (Paardekooper & Mellema 2006b) to solve the result-
ing set of equations. The computational domain extends from
r/rp = 0.4 to r/rp = 2.5, and covers the full 2π in azimuth. This
domain is covered with a regular grid with 256 cells in the radial
and 768 cells in the azimuthal direction. Nonreflecting bound-
ary conditions were used. We vary the planet mass, Mp, and f
to study different regimes of disc-planet interactions in subkep-
lerian discs.
3. Results
The most important difference between subkeplerian discs and
Keplerian discs, as far as disc-planet interactions are concerned,
is the change in position of the Lindblad and corotation reso-
nances. It was shown in Goldreich & Tremaine (1979) that an
embedded object can exchange angular momentum with the disc
only at the location of a resonance. Corotation resonances oc-
cur where the disc corotates with the planet, and inner (outer)
Lindblad resonances are located inside (outside) corotation. The
strongest Lindblad resonances occur at a distance ∼ H of the
planet Ward (1997). In Keplerian discs, a planet on a circular or-
bit is located close to corotation, with Lindblad resonances on ei-
ther side of its orbit. In strongly subkeplerian discs, the distance
of the planet to corotation is large enough so that we can safely
neglect any nonbarotropic effects associated with the corotation
1
0.0
0.1
h=0.05
h=0.1
h=0.2
0.2
1-f
0.3
0.4
Fig. 2. Absolute value of the torque on a q = 1.26 · 10−5 planet
embedded in discs of constant initial surface density for different
values of h and f . Solid curves denote the results of linear calcu-
lations, while the symbols denote results from hydrodynamical
simulations.
torque (Paardekooper & Papaloizou 2008). In this sense, subke-
plerian discs are easier to handle, since we only have to deal with
Lindblad resonances.
The approximate position of the Lindblad resonances up to
10th order are shown in Fig. 1 for different values of f . Pressure
effects have been ignored in calculating these positions. In a
purely Keplerian disc ( f = 1), the planet interacts with both
inner and outer resonances. For f < 1, the resonance positions
move inward, with the result that the planet now mainly interacts
with the outer Lindblad resonances. This has consequences for
both migration and gap formation.
3.1. Migrationoflow-massplanets
We first consider the torque Γ on low-mass planets, not massive
enough to open up a gap. We compare results from linear calcu-
lations to hydrodynamical simulations for different values of f
in Fig. 2. The torque is normalised by Γ0 = q2Σpr4
p/h2, where
q = Mp/M∗ and Ωp is the planet's angular velocity. We plot the
absolute value of the torque; it is negative for all values of f .
pΩ2
It is immediately clear from Fig. 2 that the torque very
strongly depends on f . This can be understood in terms of a
change in the position of the resonances with respect to the
planet. For f = 1 we find the classical Type I torque, modified
by nonlinear effects at corotation (Paardekooper & Papaloizou
2009). Around 1 − f ≈ h, the planet mainly interacts with the
strongest outer Lindblad resonances, yielding a very strong neg-
ative torque. The total torque approaches the one-sided Lindblad
torque (Ward 1997), which would be obtained when considering
only outer Lindblad resonances and which is a factor of ∼ 1/h
stronger than the differential Lindblad torque. The maximum
torque depends on the adopted smoothing, with higher values for
lower values of b. For 1 − f ≫ h, the torque decreases, since the
resonances it interacts with become weaker as the planet moves
S.-J. Paardekooper: Disc-planet interactions in subkeplerian discs
3
further away from corotation. For h = 0.05 and f = 0.6, the re-
sulting torque is comparable in magnitude to the classical Type I
torque. Linear calculations and nonlinear simulations agree very
well, which again indicates that nonlinear effects at corotation
do not play an important role.
The results depicted in Fig. 2 indicate that the crucial param-
eter determining the total torque on the planet is p = (1 − f )/h.
The maxima of the curves occur factors of 2 apart in Fig. 2, con-
firming this picture. The scaling of the torque with Γ0 ensures
that for f = 1, the curves fall on top of each other. When p ≈ 1,
a sizeable fraction of the one-sided Lindblad torque can be ex-
erted on the planet. The one-sided torque scales as h−3, while the
differential Lindblad torque scales as h−2. This is the reason for
the different maxima for different values of h. This scaling holds
for all f < 1; only for f = 1 do we get the h−2 scaling of the dif-
ferential Lindblad torque. Therefore, in subkeplerian discs, for a
fixed value of p, the torque scales as h−3. For a fixed value of f ,
the situation can be different. For example, for 1 − f = 0.4, the
torque for h = 0.2 is twice as strong as would be expected from
an h−3 scaling, which is due to the shift of the maximum torque
to higher values of f .
The torque due to a headwind in a subkeplerian disc on a
planet with radius Rp is given by Adams et al. (2009):
ΓX /Γ0 =
CD(1 − f )2 h
rp !2
q2 Rp
π
2
,
(1)
with CD a constant of order unity (Weidenschilling 1977). This
means that the ratio of the torque due to a headwind and the
classical Type I torque is proportional to h, and is of order unity
for an Earth mass planet located at 1 AU, in a disc with f = 0.66
and h = 0.05. Since most resonances are located far inward with
respect to the orbit of the planet, the actual Type I torque is close
to its classical value at f = 1 (see Fig. 2). It is however very
difficult to generalise this result due to the complex behaviour of
the Type I torque when varying h and f . For example, while for
f = 0.6, ΓX ≈ Γ at 1 AU for a planet of 1 M⊕ in a disc with h =
0.05, the Type I torque will in fact dominate for higher values of
h. The dimension of the parameter space is quite high (q, rp, h,
etc.), so that it is difficult to make more general statements.
It is important to realise that subkeplerian discs are haz-
ardous environments for low-mass planets. Since the planet
mainly interacts with outer Lindblad resonances, and corotation
torques play a minor role, migration will be directed inward. The
time scale for migration can be even smaller than the classical
Type I time scale, in unfavourable cases more than an order of
magnitude smaller.
3.2. Gapformation
We now turn to the issue of gap formation. The presence of a gap
determines whether a planet moves in the Type I or the Type II
regime, the latter being a much slower mode of migration (Ward
1997). There are two requirements that have to be fulfilled to
open a gap: angular momentum transport due to the presence of
the planet must exceed the viscous transport, otherwise the gap
could be refilled by viscous evolution of the disc, and, in discs
with low viscosity, the waves transporting angular momentum
should be damped locally. For standard parameters, both criteria
give similar minimum masses for gap opening, of the order of
100 M⊕. A unified criterion is derived in Crida et al. (2006).
Adams et al. (2009) derived a minimum mass for gap-
= π(1 − f )2αh2. They noted
opening in subkeplerian discs, q5/3
min
1.5
1.0
0
Σ
Σ
/
0.5
0.0
0.5
1.0
f=1
f=0.8
f=0.6
f=0.4
r0=rp
r0=rp+H
r0=rp+2H
0.6
f
0.8
1.0
2.0
)
0
r
(
Σ
1.0
0.8
0.6
0.4
0.2
0.0
0.4
1.5
r/rp
Fig. 3. Azimuthally averaged surface density, in units of the ini-
tial (constant) surface density Σ0, for a q = 5 · 10−4 planet em-
bedded in a h = 0.05, α = 0.001 disc for different values of f .
The inset shows the averaged surface density at the orbital radius
of the planet as a function of f .
that for f = 0.66 this mass is comparable to the classical esti-
mates for Keplerian discs, although the planet would be located
more towards the inner edge of the gap. In this section, we study
gap formation for a q = 5 · 10−4 (corresponding to 0.5 MJ around
a 1 M⊙ star) planet embedded in a disc with h = 0.05 and a
viscosity corresponding to α = 0.001. For 0.4 ≤ f ≤ 1, this
planet fulfils the criterion for gap formation as given above, if
only slightly for f = 0.4.
The results are depicted in Fig. 3, where we show the az-
imuthally averaged surface density for different values of f af-
ter 500 orbits of the planet. The system has reached a steady
state by that time. For a Keplerian disc, the surface density close
to the planet has dropped by more than an order of magnitude,
confirming that this planet indeed opens up a gap. The main dif-
ference between Keplerian and subkeplerian discs is the posi-
tion of the gap, which is shifting inward for f < 1. This can
again be understood in terms of resonance positions: the impor-
tant resonances are located inside the orbit of the planet. The
associated waves are damped locally through shocks, which re-
sults in a gap at r < rp. Therefore, contrary to the findings of
Adams et al. (2009), the planet is located near the outside of the
gap. The analysis of Adams et al. (2009) only applies to the case
1 − f < h, for which the location of the original gap (for f = 1)
is still densely populated by resonances. In this case, because the
outer resonances are located closer to the planet (but still outside
its orbit), the outer part of the gap forms more easily, leaving the
planet closer to the inner edge of the gap.
The depth of the gap does not strongly depend on f . As long
as 1 − f is not larger than the width of the gap for f = 1, the
angular momentum flux will be similar in both cases. Then, the
ability of a planet to open up a gap is relatively independent of
f . However, from the perspective of the planet the case f < 1
can be radically different from the Keplerian case. For what ac-
cretion and migration are concerned, what matters is the surface
S.-J. Paardekooper: Disc-planet interactions in subkeplerian discs
4
0
Σ
Σ
/
1.2
1.0
0.8
0.6
0.4
0.2
0.0
f=1
f=0.8
f=0.6
f=0.4
0.5
1.0
1.5
r/rp
2.0
for Keplerian discs that magnetic fields can have a strong
impact on planet migration: regular fields introduce mag-
netic resonances (Terquem 2003), while magnetic turbulence
introduces stochastic migration (Nelson & Papaloizou 2004;
Adams & Bloch 2009). It remains to be seen what impact a mag-
netic field configuration that gives rise to subkeplerian discs can
have on the simple hydrodynamic picture presented here.
We have worked in the isothermal limit, but since corota-
tion torques only play a minor role in strongly subkeplerian
discs, results for more realistic discs should be similar. The
two-dimensional approximation, in combination with a grav-
itational softening parameter of order h, gives similar results
to fully three-dimensional simulations, again at least as far as
the Lindblad torque is concerned (Paardekooper & Papaloizou
2009). We have found that the results do not depend strongly on
the initial surface density profile.
We have considered migration of low-mass planets (the Type
I regime), finding that there is a strong dependence on f . For 1 −
f ≈ h, the planet feels almost the full one-sided Lindblad torque,
which is a factor 1/h stronger than the classical Type I torque.
Such a disc would be very hazardous to low-mass planets, since
inward migration is sped up by more than an order of magnitude
compared to Keplerian discs. For 1 − f > h, the torque decreases
because the resonances the planet interacts with become weaker.
The dependence of the torque on h and f is quite complicated,
and it is not easy to say for a disc of given f and h whether
the Type I torque will be stronger of weaker than the head-wind
torque.
Gap formation proceeds similar to that in Keplerian discs.
However, because of the inward shift of the important reso-
nances, the gap will be located inside the planet's orbit. This then
leaves the planet on the outside of its own gap. For strongly sub-
keplerian discs, a gap-opening planet can become fully embed-
ded again. Since the torque-generating resonances are located
far away, this does not restore the full Type I torque. Accretion
time scales could be very short, if the planet is able to accept the
available matter.
I acknowledge support from STFC in the form of a post-
Acknowledgements.
doctoral fellowship. I wish to thank Pawel Ivanov for his interest in retrograde
orbiting embedded planets, which formed the start of this project, and the anony-
mous referee for an insightful report.
References
Adams, F. C. & Bloch, A. M. 2009, ApJ, 701, 1381
Adams, F. C., Cai, M. J., & Lizano, S. 2009, ApJ, 702, L182
Bate, M. R., Lubow, S. H., Ogilvie, G. I., & Miller, K. A. 2003, MNRAS, 341,
213
Crida, A., Morbidelli, A., & Masset, F. 2006, Icarus, 181, 587
D'Angelo, G., Kley, W., & Henning, T. 2003, ApJ, 586, 540
Goldreich, P. & Tremaine, S. 1979, ApJ, 233, 857
Lin, D. N. C. & Papaloizou, J. 1986, ApJ, 309, 846
Masset, F. S. & Papaloizou, J. C. B. 2003, ApJ, 588, 494
Nelson, R. P. & Papaloizou, J. C. B. 2004, MNRAS, 350, 849
Paardekooper, S.-J. & Mellema, G. 2006a, A&A, 459, L17
Paardekooper, S.-J. & Mellema, G. 2006b, A&A, 450, 1203
Paardekooper, S.-J. & Mellema, G. 2008, A&A, 478, 245
Paardekooper, S.-J. & Papaloizou, J. C. B. 2008, A&A, 485, 877
Paardekooper, S.-J. & Papaloizou, J. C. B. 2009, MNRAS, 394, 2283
Papaloizou, J. C. B., Nelson, R. P., Kley, W., Masset, F. S., & Artymowicz, P.
2007, in Protostars and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil, 655 --
668
Pepli´nski, A., Artymowicz, P., & Mellema, G. 2008, MNRAS, 386, 179
Shu, F. H., Galli, D., Lizano, S., Glassgold, A. E., & Diamond, P. H. 2007, ApJ,
Fig. 4. Azimuthally averaged surface density, in units of the ini-
tial (constant) surface density Σ0, for a q = 10−4 planet embed-
ded in a h = 0.05, α = 0.004 disc for different values of f .
density close to the planet. Since the planet is located more and
more outside the gap for low values of f , it could accrete and
migrate as if there were no gap. This is illustrated in the inset of
Fig. 3, where we show the azimuthally averaged surface density
at three different locations in the disc for different values of f . In
Keplerian discs, planets usually interact with material within 2
scale heights from their orbit (Bate et al. 2003). This is the max-
imum distance that is considered in the inset of Fig. 3.
For a Keplerian disc, a gap forms with a half width that is
approximately 2H, and in the inset of Fig. 3 all three curves fall
below 0.1 for f = 1. For f = 0.8, there is still a density de-
pression around the orbit of the planet, but for lower values of f
there is no clear gap near r = rp, it has shifted inwards enough so
that the planet is basically embedded in the disc again. However,
since for the lowest values of f there are no more resonances
located close to the planet, the Type I torque will not be fully
restored. We find that typically Γ( f = 0.6)/Γ( f = 1) ≈ 2 for
gap-opening planets. The influence of the head wind (see Eq. 1)
will be strong, however, since it does not rely on resonances. We
comment that a gap-opening planet is still tidally locked to the
gap, just as in the Keplerian Type II migration case.
The possibility of a high-mass planet fully embedded in the
disc may have some important consequences for gas accretion.
In a Keplerian disc, accretion drops by an order of magnitude
when a gap is formed (D'Angelo et al. 2003). In a subkeplerian
disc, there a significant amount of mass remains near the orbit of
the planet, making accretion potentially very efficient. It is not
clear, however, if the planet is able to accept material that has
such a high relative velocity.
The results for a more massive planet of q = 0.001 (1 MJ
around a 1 M⊙ star) are very similar to those presented in Fig. 3.
The gap shifts inward, and for f < 0.7 the planet is located on
the outer edge of its own density depression. For smaller plan-
ets, which only open up a shallow density depression for f = 1,
remain fully embedded for f < 1, as shown in Fig. 4. While for
f = 1, a q = 10−4 planet decreases the surface density around its
orbit by a factor 0.7, for f < 0.7 there is no more evidence for
any density depression. For this lower planet mass, the impor-
tant resonances become too weak to affect the surface density in
strongly subkeplerian discs.
4. Discussion and conclusions
We have presented hydrodynamical simulations of planets em-
bedded in subkeplerian discs. They represent the first step
towards modelling fully 3D magnetised discs. It is known
665, 535
Terquem, C. E. J. M. L. J. 2003, MNRAS, 341, 1157
Ward, W. R. 1997, Icarus, 126, 261
Weidenschilling, S. J. 1977, MNRAS, 180, 57
|
1511.00009 | 1 | 1511 | 2015-10-30T20:00:02 | Oscillations of Relative Inclination Angles in Compact Extrasolar Planetary Systems | [
"astro-ph.EP"
] | The Kepler Mission has detected dozens of compact planetary systems with more than four transiting planets. This sample provides a collection of close-packed planetary systems with relatively little spread in the inclination angles of the inferred orbits. A large fraction of the observational sample contains limited multiplicity, begging the question whether there is a true diversity of multi transiting systems, or if some systems merely possess high mutual inclinations, allowing them to appear as single-transiting systems in a transit-based survey. This paper begins an exploration of the effectiveness of dynamical mechanisms in exciting orbital inclination within exoplanetary systems of this class. For these tightly packed systems, we determine that the orbital inclination angles are not spread out appreciably through self-excitation. In contrast, the two Kepler multi-planet systems with additional non-transiting planets are susceptible to oscillations of their inclination angles, which means their currently observed configurations could be due to planet-planet interactions alone. We also provide constraints and predictions for the expected transit duration variations (TDVs) for each planet. In these multi-planet compact Kepler systems, oscillations of their inclination angles are remarkably hard to excite; as a result, they tend to remain continually mutually transiting (CMT-stable). We study this issue further by augmenting the planet masses and determining the enhancement factor required for oscillations to move the systems out of transit. The oscillations of inclination found here inform the recently suggested dichotomy in the sample of solar systems observed by Kepler. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 1–14 (2015)
Printed 10 July 2018
(MN LATEX style file v2.2)
Oscillations of Relative Inclination Angles
in Compact Extrasolar Planetary Systems
Juliette C. Becker1 and Fred C. Adams1,2
1Astronomy Department, University of Michigan, Ann Arbor, MI 48109
2Physics Department, University of Michigan, Ann Arbor, MI 48109
February 2015
ABSTRACT
The Kepler Mission has detected dozens of compact planetary systems with more
than four transiting planets. This sample provides a collection of close-packed plane-
tary systems with relatively little spread in the inclination angles of the inferred orbits.
A large fraction of the observational sample contains limited multiplicity, begging the
question whether there is a true diversity of multi-transiting systems, or if some systems
merely possess high mutual inclinations, allowing them to appear as single-transiting
systems in a transit-based survey. This paper begins an exploration of the effectiveness
of dynamical mechanisms in exciting orbital inclination within exoplanetary systems
of this class. For these tightly packed systems, we determine that the orbital inclina-
tion angles are not spread out appreciably through self-excitation. In contrast, the two
Kepler multi-planet systems with additional non-transiting planets are susceptible to
oscillations of their inclination angles, which means their currently observed config-
urations could be due to planet-planet interactions alone. We also provide constraints
and predictions for the expected transit duration variations (TDVs) for each planet. In
these multi-planet compact Kepler systems, oscillations of their inclination angles are
remarkably hard to excite; as a result, they tend to remain continually mutually tran-
siting (CMT-stable). We study this issue further by augmenting the planet masses and
determining the enhancement factor required for oscillations to move the systems out
of transit. The oscillations of inclination found here inform the recently suggested di-
chotomy in the sample of solar systems observed by Kepler.
Key words: planets and satellites: dynamical evolution and stability - planetary sys-
tems
1
INTRODUCTION
The Kepler mission has discovered a large number of compact
extrasolar systems containing multiple planets that can be ob-
served in transit (Lissauer et al. 2011a; Batalha et al. 2013).
Roughly forty of these such systems have four or more planets.
The inventory of these four-plus planet systems includes mostly
super-Earth sized planets, which have radii RP = 2−5R⊕ and
orbital periods in the range 1 – 100 d. Moreover, the orbital pe-
riods of the planets within a given system are regularly spaced
(roughly logarithmically uniform in period or semimajor axis).
Because all of the planets were observable by Kepler at their
times of discovery, these systems have an additional stringent
dynamical constraint: they must have retained a relatively nar-
row spread in their orbital inclination angles. On the other hand,
orbital inclination can often be excited in close-packed plane-
tary systems. The goal of this paper is thus to explore the oscil-
c(cid:13) 2015 RAS
lations of orbital inclination within solar systems of this class.
Excitation of inclination can be driven by a variety of mechan-
ims, incluing unseen additional companions, perturbations from
stellar encounters in clusters (Adams & Laughlin 2001; Li &
Adams 2015), and self-excitation through interactions among
the observed planets. This paper focuses on this latter mecha-
nism.
Slight deviations from true coplanarity in these systems
(e.g., as observationally supported in Rowe et al. 2014; Lissauer
et al. 2011b; and others) allow for the possibility of oscillations
in the inclination angles of the planetary orbits, e.g., due to secu-
lar interactions between the planets (see also Van Laerhoven &
Greenberg 2012). If such oscillations were common, and had
sufficient amplitude, then not all members of a solar system
could be seen in transit at every epoch. As a result, multi-planet
systems would display evidence for "missing" planets, i.e., ex-
ceptions to the (roughly) logarithmically even spacing of orbits
5
1
0
2
t
c
O
0
3
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
0
0
0
0
.
1
1
5
1
:
v
i
X
r
a
(cid:32)
(cid:19)1.3
(cid:18) R
R⊕
(cid:33)
2
Becker & Adams
that are often observed. The ubiquity of this class of exoplane-
tary systems places strong constraints on both their architectures
and dynamical histories (see also Chiang & Laughlin 2013). We
note that the inclination angle oscillations for Jupiter and Saturn
in our own solar system are large enough to periodically move
the orbits out of a mutually transiting configuration.
Statistical analyses of the Kepler system architectures sug-
gest that there could exist two distinct populations of plane-
tary systems (Ballard & Johnson 2014; Morton & Winn 2014),
namely, a population with single-transiting planets and an ad-
ditional population of multi-planet systems. The existence of
these two distinct populations could be explained by either
two true distributions of solar systems (e.g., created by two
different formation histories) or a single distribution in which
some systems exhibit a high degree of scatter in orbital incli-
nation angles. Excitation of inclination in nearly coplanar sys-
tems could shift some planets out of a transiting configuration,
thereby leading to the population of single-transit systems. In
this case, the single-transit systems would be a subset of the
multi-transiting group rather than a distinct population.
This paper explores possible oscillations of the inclination
angles in compact extrasolar systems. The measured planetary
radii RP = 2 − 5R⊕ imply planetary masses MP = 4 −
30M⊕, where we use a conversion law based primarily on the
probabilistic mass-radius relationship derived in Wolfgang et al.
(2015):
∼ Normal
M
M⊕
µ = 2.7
, σ = 1.9
(1)
where M refers to the mass of a body, R its radius, and this ex-
pression represents a r1.3 scaling law with a normal distribution
of scatter due to potential planetary composition variation. The
Wolfgang relationship describes the a distribution of the poten-
tial masses for planets in the range RP = 1.5 − 4R⊕. Since a
small number of planets in our sample lie outside these bounds,
we supplement the Wolfgang relation in two ways: for planets
with radii RP < 1.5R⊕, we supplement with the rocky relation
from Weiss & Marcy (2014); for planets with radii RP > 4R⊕,
we determine starting density using the Wolfgang relation, then
add a scatter and choose a radius anomaly to account for vary-
ing core masses and inflation due to thermal effects (Laughlin
et al. 2011). Of the 208 planets in our sample, only 9 fall above
the regime described by the Wolfgang relation. With relatively
large masses and close proximity, planet-planet interactions can
be significant. On the other hand, these planetary systems orbit
relatively old stars (with ages of ∼ 1 − 6 Gyr, weighted toward
the lower end of this range; see Walkowicz & Basri 2013), so
that they are expected to be dynamically stable over ∼Gyr time
scales. These systems are also generally non-resonant. These
considerations - significant interactions coupled with long-
term stability and non-resonance - suggest that the planetary
systems are subject to secular interactions (Murray & Dermott
1999). In the present context, we are interested in secular oscil-
lations of the inclination angles of the orbits. If such oscillations
have sufficient amplitudes, the resulting spread of inclinations
angles in the system will sometimes be large enough that not
all of the planets can be seen in transit. When observed in such
a configuration, the system will appear to have gaps in the reg-
ular spacing of planetary orbits that these systems usually ex-
hibit. The goal of this paper is to understand the amplitude of
self-excitation of inclination angle oscillations and provide lim-
its on transit duration variations, an observable with amplitude
directly related to inclination evolution over time, for observed
Kepler systems with no unseen companions. This analysis will
allow future observations of transit durations for these systems
to inform the presence of massive outer companions in these
systems.
We note that spreads in the inclination angles can be pro-
duced by a variety of astronomical processes. This work will
focus on secular oscillations of the inclination angles by the
compact solar system planets themselves (with semi-major axes
a <∼ 0.5 AU). Future work will focus on the effect of possible
additional bodies in the outer part of the solar system (where
a ≈ 5 − 30 AU), roughly analogous to the giant planets in our
outer Solar System.
We stress that oscillations of inclination angles are not rare.
Within our Solar System, for example, the orbital inclinations of
Jupiter and Saturn oscillate with a period of about 51,000 years
and an amplitude of about 1◦ (see Figure 7.1 in Murray & Der-
mott 1999). The inclination angles of the two orbits coincide
every half period (25,500 years), so that an observer oriented in
that plane would see both planets in transit at that epoch. How-
ever, the amplitude of the oscillation is sufficient to move both
planets out of transit for an appreciable fraction of the secular
cycle.
This paper focuses on the case of self-excitation of inclina-
tion angles for Kepler systems with four or more planets, where
the secular dynamics of such systems are considered in Section
3. An analysis of the observed compact, mutually transiting sys-
tems is presented in Section 3.1, which shows that the systems
are consistently mutually transiting over time. An orbital archi-
tecture that is continually mutually transiting is denoted here as
CMT-stable (which should not be confused with dynamical sta-
bility). We consider a generalized class of systems in Section
3.2, and study compact systems which have been discovered to
host an additional non-transiting planet in Section 3.3 (where
these systems are shown to be more active). We also compare
these results with numerical simulations in Section 3.4. Section
4 presents observables for the compact Kepler systems discov-
ered to date; specifically, the transit durations are predicted to
vary and the magnitude of these variations are determined. In
Section 5, we study the stability of the observed Kepler systems
by considering how the predicted oscillation amplitudes would
vary if planet masses are scaled upward: the systems are found
to be remarkably dynamically stable. The paper concludes, in
Section 6, with a summary of our results and a discussion of
their implications, as well as a statement on our plans for future
work.
2 SECULAR THEORY FOR INCLINATION ANGLES
To evaluate the behavior of mutual inclination for these isolated
systems, we apply Laplace–Langrange secular theory (Murray
& Dermott 1999). This formalism allows the use of the long-
period terms of the disturbing function to describe orbital mo-
tion over many secular periods.
2.1 Review of the Theory
We expand to second order in inclination and eccentricity, and
first order in mass. With this expansion, inclination and eccen-
tricity are decoupled, so we can write the disturbing function as
c(cid:13) 2015 RAS, MNRAS 000, 1–14
(cid:34)
a function of inclination alone:
R(sec)
j
= nja2
j
1
2
BjjI 2
j +
(BjkIjIk cos (Ωj − Ωk))
(cid:35)
,
N(cid:88)
k=1
(2)
where j is the planet number, n is the mean anomaly, I is the
inclination, ω is argument of pericenter, and Ω is the longitude
of the ascending node. The coefficients Bij are defined by
(cid:18) Rc
aj
(cid:19)4 − 15
J 2
4
4
J 2
2
(cid:19)4
(cid:18) Rc
(cid:35)
aj
+
Mc + mj
αjk ¯αjkb(1)
3/2(αjk)
,
(3)
Bjj = −nj
and
(cid:34)
3
2
J2
8
aj
(cid:19)2 − 27
(cid:18) Rc
(cid:88) mk
(cid:20) 1
1
4
mk
4
Mc + mj
Bjk = nj
αjk ¯αjkb(1)
3/2(αjk)
(cid:90) 2π
0
where J2 and J4 describe the oblateness of the central star
(which we set to be = 0 in all our analysis), Rc is the stellar
radius, mk indicates the mass of the kth planet, Mc denotes the
mass of the central star, αjk denotes the semi-major axis ratio
aj/ak, and ¯αjk denotes the semi-major axis ratio aj/ak < 1.
The quantities b(1)
3/2 is the Laplace coefficient, which is defined
by
b(1)
3/2 =
1
π
cos ψ dψ
(1 − 2α cos ψ + α2)3/2 ,
(5)
(as given in Murray & Dermott 1999). All of the coefficients
Bjk can be considered as frequencies that describe the interac-
tion between each pair of planets, and are elements of the ma-
trix denoted as B. This application of secular theory allows us
to evaluate the problem analytically, but neglects higher-order
terms. In this formulation, the only terms in the disturbing func-
tion are those that do not depend on the mean longitudes, as
we assume that the short-period terms average out over long
timescales. The coefficient matrix B describes inclination evolu-
tion. Solving for the matrix elements of B allow us to determine
the time evolution of inclination.
The matrix B defines an eigenvalue problem (Murray &
Dermott 1999), where the eigenvalues describe the interaction
frequencies between any pair of planets. The eigenfrequencies
of this matrix, denoted here as fi, along with the eigenvectors
Ijk, can be used to describe the time evolution of the system.
Given the matrix B, we can solve for the eigenvalues and eigen-
vectors using standard methods. With these quantities specified,
we also need the initial conditions to specify the full solution
for the time evolution of the inclination angles Ij and the an-
gles Ωj. It is convenient to transform the dependent variables
according to
pj = Ij sin Ωj
and
qj = Ij sin Ωj ,
so that the solutions take the form
N(cid:88)
N(cid:88)
k=1
pj(t) =
Ijk sin(fkt + γk)
and
qj(t) =
k=1
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Ijk cos(fkt + γk) ,
(6)
(7)
(8)
Inclination Oscillations for Exoplanet Orbits
3
where the phases γk, along with the overall amplitudes, are de-
termined by the initial conditions. The quantities Ijk are eigen-
vectors, where we use the standard (but awkward) notation such
that the first index j specifies the planet number and hence the
components of the eigenvector and the second index k runs over
the different eigenvectors. It is also useful to define normalized
eigenvectors Ijk and corresponding scaling factors Tk such that
(9)
Ijk = TkIjk .
The initial conditions then specify the scaling factors through
the expressions
pj(t = 0) =
TkIjk sin γk
(cid:21)
and
,
(4)
qj(t = 0) =
TkIjk cos γk .
(10)
(11)
N(cid:88)
N(cid:88)
k=1
k=1
(cid:113)
The scaled eigenvectors Ijk (which conform to the sys-
tem's boundary conditions), the eigenvalues fk, and the phases
γk are sufficient to specify the time evolution of the orbital in-
clination of each body in the system, i.e.,
Ij(t) =
[pj(t)]2 + [qj(t)]2 ,
(12)
where the solutions pj(t) and qj(t) are given by equations (7)
and (8). Implicit in this solution is the linear dependance on the
interaction coefficients (the matrix elements given by equations
[4]). From this solution, we note that the inclination evolution
has a linear dependance on mass ratio and a second order de-
pendence on the semi-major axis ratio between the planet in
question and each planet exterior to its orbit.
3 INCLINATION OSCILLATIONS DUE TO
SELF-EXCITATION
The compact Kepler systems with four or more planets are
tightly packed systems with minimal mutual inclinations. From
this population, it appears that planets in multi-transiting sys-
tems generally have non-null values of mutual impact param-
eter, and subsequently inclination (Rowe et al. 2014). Systems
with non-null mutual inclinations exhibit non-parallel angular
momentum vectors, allowing the possibility of excitation in in-
clination and other orbital elements. To test the magnitude of
these excitations, we take the population of all Kepler systems
with four or more transiting planets as examples of compact,
multi-body, transiting systems. We obtain our data from the
NASA Exoplanet Archive1, updating system parameters when
newer values have been found (such as in the case of Kepler-
296; Barclay et al. 2015).
There are observational biases inherent in the Kepler sys-
tems, as a photometric transit survey is by definition more likely
to find systems with low mutual inclinations and aligned argu-
ment of pericenters (Ragozzine & Holman 2010). The Kepler
multi-planet systems are likely more aligned and more com-
pact in inclination plane width than an 'average' system, but
the sample found by Kepler is representative of the type of sys-
tem we would expect to see from photometric transit surveys
1 http://exoplanetarchive.ipac.caltech.edu
4
Becker & Adams
Figure 1. Plotted here are the inclination evolutions of five roughly coplanar planets, with initial conditions drawn from the priors of Kepler-256.
Although the inclinations of the planets generally stay within a plane, there is also instantaneous variation, which manifests as a range in the inclinations
of plane of planets. This variation may lead planets to be knocked out of a transiting configuration. The mutual inclination, shown on the right panel,
changes as planets precess, meaning that the width of the plane containing all the planets oscillates over time.
such as Kepler (Borucki et al. 2011), K2 (Howell et al. 2014),
and TESS (Ricker et al. 2014). It is not currently clear, however,
whether the Kepler multi-planet systems that we do see in tran-
sit are CMT-stable or if we are catching them at a lucky moment
in which all planets appear to be in transit. This differentiation
is important because the former possibility describes a much
less dynamically active system than the latter. To test the stabil-
ity against exciting planets out of the transiting plane, we used
the secular theory described in Section 2 to numerically evolve
each system in the Kepler multi-planet sample for several secu-
lar periods. This procedure results in a measure of the spread in
impact parameter ∆b(t) (see below). We also compute the prob-
ability that the system is mutually transiting, marginalized over
all trials and realizations in our simulations. If ∆b(t) < 2 for an
entire secular period and the probability of all planets transiting
simultaneously for a random time-step in a random realization
of the system is high (P (transit) > 0.85) then the system is
said to be CMT-stable in a transiting configuration. Note that
the condition of being CMT-stable against oscillating planets
out of transit is much more confining that being dynamically
stable against planet ejection. For a given Kepler system, we can
use a Monte Carlo analysis to evaluate ∆b(t) not just once, but
many times, with starting orbital elements for each realization
selected from observationally motivated priors. For parameters
that have been measured (for transiting systems, the radius of
the planet rp and the semi-major axis ap, and sometimes the
inclination Ip or eccentricity ep), we draw each planet's orbital
element from a normal distribution with mean and standard de-
viation determined from observations. For orbital elements not
measured, we draw a value from priors summarized in Table 1.
Observationally measured inclinations have been fit from
photometric light curves, and for these planets there is a de-
generacy between angles over 90◦and under 90◦. The literature
reports inclination angles as < 90◦, so when we use a litera-
ture measurement, we choose a value not only from that planet's
measured posterior but also choose its orbit to fall above or be-
low the midplane of the star with equal probability. For planets
without measured inclinations, we choose a plane width from
a Rayleigh distribution with width 1.5◦(Fabrycky et al. 2014),
subject to the constraint that all planets must be transiting. This
choice of distribution follows work done by Fabrycky & Winn
(2009); Lissauer et al. (2011a); Fang & Margot (2012); Ballard
Orbital Element Distributions
Parameter
Prior
ω
Ω
e
I
uniform on (0◦, 360◦)
uniform on (0◦, 360◦)
uniform on (0, 0.1)
Rayleigh distribution with width σ = 1.5◦
Table 1. When orbital elements have not been measured observation-
ally, we draw their values randomly from the prior distributions sum-
marized in this table.
& Johnson (2014). In these recent works, Rayleigh distributions
with varying widths are used to describe the size of the plane
containing the planets. The value we use here, 1.5◦, is within
the range suggested by the work of Fabrycky et al. (2014).
We note that the argument of the ascending node is not
necessarily independent of the value of inclination angle as as-
sumed here. As planetary systems evolve to attain nonzero in-
clination angles, modeled here by a Rayleigh distribution, the
nodes will evolve into some other distribution, which should be
characterized in future work.
Once we have the initial conditions for each Kepler system,
we can evolve orbits as according to the secular theory described
in Section 2. This must be done individually for each realization
of initial conditions for each system.
3.1 Evaluating the Secular Behavior of the Compact,
Multi-Planet Kepler Systems
A tightly packed, roughly coplanar system of planets will trade
angular momentum as the system evolves (while keeping the
total angular momentum vector of the system constant). The
magnitude of this exchange determines the magnitude of the
variations in orbital elements of each body. Equation (12) de-
scribes the inclination evolution for each body in a system. Once
the inclination solutions for each planet in a system have been
found using equation (12), a comparison between them (see Fig-
ure 1, which illustrates how the mutual inclination can change
over time) yields a measure of the mutual inclination between
c(cid:13) 2015 RAS, MNRAS 000, 1–14
range in inclinationInclination Oscillations for Exoplanet Orbits
5
Figure 2. The parameterization of mutual impact parameter, as illus-
trated by test case Kepler-11. First, a plot of inclination for all bodies in
a system (upper left) is generated by solving equation (12) with the ini-
tial parameters of the system as boundary conditions. The semi-major
axis dependency is removed using equation (13), and the result, impact
parameter over time for each planet, is shown (upper right). The inclina-
tions attained by each planet result in vastly different impact parameters
due to the differences in semi-major axis. Planets closer to the star can
attain more inclination with less effect on their impact parameter. Fi-
nally, the range between impact parameters is calculated (lower right)
as was done for mutual inclination in Figure 1. The result is a measure
of the range of the mutual impact parameter over time, ∆b(t). As long
as this width describes a plane that lies entirely within the limbs of the
star, the planets will be CMT-stable.
all planets in the system. This mutual inclination describes the
width of the plane containing all the planets.
As the condition for transiting is more rigorous than ap-
proximate coplanarity (even as planets' inclinations vary in con-
cert, the planets with larger orbital separations are more likely
to cease transiting), we remove the dependence on orbital and
stellar properties by working in terms of impact parameter, b,
which is defined as:
bj =
aj
R∗
cos (Ij)
(13)
where j is planet number, a is the semi major axis, R∗ the radius
of the central star, and I the inclination. When −1 < bj < 1,
planet j will transit. Using the analytic expression for inclina-
tion evolution (Equation 12), we can describe the long-term be-
havior of not only individual planets but the range of their re-
spective impact parameters. The process of extracting the mu-
tual impact parameter ∆b is shown in Figure 2.
Using this technique, we explored the evolution of orbits
for the entire initial condition parameter space for each Kepler
multi-planet system. For a given system, we conducted 4000
Monte Carlo trials for each Kepler system, resulting in 4000 re-
alizations of ∆b(t), with different initial conditions drawn from
the observational priors, supplemented with the values in Table
1. This sample can be used to calculate the mean range of the
impact parameter over time for the Kepler system, as well as the
width of the plane of planets in impact-parameter space.
We repeated this process of 4000 Monte Carlo trials for
each for the 43 systems in our sample of multi-planet Kepler
systems, resulting in a measure of the inclination evolution be-
havior for each system. Figure 3 visualizes the results of these
trials, where each point represents the mean mutual impact pa-
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Figure 3. For each multi-planet Kepler system, the parameters of the
system were sampled 4000 times and evolved forward in time. The re-
sulting inclination angles for the planetary orbits were converted to a
mutual impact parameter (see text). The mean and scatter of these val-
ues are plotted here for each system as a function of the total mass of the
transiting planets, given in earth masses. The dotted horizontal line in-
dicates the level above which it is not possible to observe all the planets
in transit.
rameter for a different Kepler system. Mean mutual impact pa-
rameter is the typical width of the plane containing all planets
in the system, and must be smaller than the diameter of the star
for all planets to transit. An impact parameter plane width of
∆b = 2, marked on the plot, is the upper limit for all planets in
the plane to be transiting.
For each point in Figure 3, the height of the point as com-
pared to the transit limit (δb = 2) corresponds to the width of
the plane containing all the planets. The scatter (represented by
error bars) corresponds to the width of the distribution due to the
variations between realizations. For all systems, the projected
plane containing the planets is much smaller than the diameter
of the star, which means we would expect to see all the planets
in transit at for the majority of the secular history of the system.
This parameterization represents the average behavior of
each system over time. The plane width demonstrates how much
range in impact parameter is normal for each architecture of
system. However, we care about the transiting behavior of each
system with respect to a single line of sight: that of the observer
(Kepler) who originally identified the planets as mutually tran-
siting. For example, it would be possible for a system's impact
parameter range to be small enough for it to be possible for all
planets to transit, but for the plane to be situated in such a lo-
cation that only some planets transit. To understand how likely
this is to happen, we plot in Figure 4 the mutual transit prob-
ability for each observed system as blue circular points. This
probability is defined as the probability that a random time-step
from a random realization, chosen from the sample of all 4000
realizations considered in the construction of Figure 3, will have
all planets transiting along the line of sight to Earth. A proba-
bility of 1 would mean that the planets never left a transiting
configuration in any time-step in any of our simulations, while
a probability of 0 means that the system was never mutually
transiting in any time-step in any realization.
Figure 4 shows that for the observed Kepler systems, all
planets are expected to be transiting more than 85% of the time.
Indeed, for most systems the probability of mutual transit is
6
Becker & Adams
Figure 4. For all realizations considered in Sections, 3.1, 3.2, and 3.3,
we plot as circles the probability that a randomly chosen time-step from
a randomly chosen realization will have all planets transiting along the
line of sight to Earth. For the observed Kepler systems, all systems
are mutually transiting more than 85% of the time. This result indi-
cates that statistically the observed Kepler systems are seen in transit an
overwhelming majority of the time. The generalized systems, plotted as
crosses, are mutually transiting a much lower fraction of the time, as are
Kepler-48 and -68, the observed currently non transiting systems.
even closer to 100%. This demonstrates that not only do we ex-
pect the Kepler multi-planet systems to have plane widths small
enough to potentially be transiting (Figure 3), the majority of the
time they should maintain these transiting configurations with
respect to our line of sight (Figure 4).
From an analysis of the results in Figure 3 and Figure 4, it
appears that while Kepler systems do excite mutual inclinations
due to their dynamical interactions with each other (as their mu-
tual impact parameters do change over time), the magnitude of
these interactions are small enough that although an initially
non-null mutual inclination exists, it remains, through the pro-
cess of secular evolution, smaller than the threshold necessary
for planets to not be observed in transit. From this, we can state
that the observed Kepler systems are generally CMT-stable.
The Kepler systems with four or more planets do not ex-
hibit sensitivity to self-excitation of inclination due to dynamic
interactions between the inner, roughly coplanar planets. This
result indicates that self-excitation (in the mode considered
here) is not a dominant mechanism in knocking planets out a
transiting plane and thereby creating tightly-packed systems in
which only some planets transit.
It it important to note that the analysis of these observed
Kepler system is limited by several factors: the measured mu-
tual inclinations will be artificially low compared random sys-
tems drawn from the true distribution of planetary architectures,
as these are systems with narrow enough ranges in inclination
to be discovered in transit in the first place; the impact param-
eters of observed systems are likely artificially low due to the
signal-to-noise bias against higher impact parameters; the devi-
ation between measured planetary arguments of pericenter will
also be artificially low (Ragozzine & Holman 2010). These sys-
tems are not a representative sample of the true distribution of
systems. As a result, the analysis presented here for the observed
Kepler systems is not an analysis of the underlying planet popu-
lation, but only of this particular class of heretofore discovered
systems.
3.2 Inclination Oscillations in Generalized Kepler
Systems
The Kepler systems that we see are observationally biased in
that they likely have unusually low mutual inclinations and
aligned arguments of pericenter (Ragozzine & Holman 2010).
As we have shown in Section 3.1, the observed Kepler systems
are remarkably CMT-stable in their transiting configurations.
We are not simply lucky to see these systems in transit, merely
viewing them at an opportune time: instead, we are seeing sys-
tems that will likely be consistently transiting over many secu-
lar timescales. The Kepler systems are indeed a special class of
system. It would also be interesting to compare their behavior
with that of generalized Kepler systems, with a wider range of
starting orbital parameters.
To construct these systems, we repeat the following pro-
cess for each Kepler system in our sample:
• Generate a compact planetary system based on the target
Kepler system. To do this, we draw each orbital parameter from
an inflated distribution, treating measured 3σ errors as the width
of our prior from which to draw orbital parameters. We convert
radii to masses using the extended Wolfgang relation.
• We evaluate the system for dynamical stability using the
Hill-radii criteria outlined in Fabrycky et al. (2014). We com-
pute the separation between two orbits (∆) in terms of their Hill
radii:
∆ = (aout − ain)/RH
(14)
when the mutual Hill radius is given by:
(cid:18) Min + Mout
(cid:19)1/3
(aout − ain)/2
3M∗
RH =
(15)
√
and for a system to be considered dynamically stable, ∆ > 2
3
and for each pair of planets, ∆in + ∆out > 18 (Fabrycky et al.
2014).
• If the system is dynamically stable according to these Hill
arguments, we evolve the system and repeat the process for an-
other set of starting parameters.
Once this process is completed for each Kepler system, we have
a sample of analog Kepler systems, which are based on the ob-
served systems but no longer exactly the systems that we ob-
serve. This sample allows us to compute the mean mutual im-
pact parameter over time, just as we did for the observed Kepler
systems in the previous section.
The result is shown in Figure 5, which shows the same
statistic plotted in Figure 3 computed from the generalized Ke-
pler systems. For these generalized systems, the range of the
impact parameter over time is higher, suggesting that the Kepler
systems we observe are a particularly CMT-stable subset of the
dynamically possible compact systems that could exist. Figure
4 shows as red crosses the mutual transit probability (over all
time-steps and all realizations) for these generalized systems,
demonstrating that the generalized systems spend significant
amounts of time in non-mutually transiting configurations, as
their plane widths imply they should.
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Inclination Oscillations for Exoplanet Orbits
7
and then through secular interactions some planets had been
perturbed out of the transiting plane.
We can test this explanation for the currently observed mis-
alignment of Kepler systems that have been discovered to have
multiple transiting planets and additional, non-transiting com-
panions using the same method that was used to evaluate the
transit stability of the most tightly packed Kepler systems in
Section 3. Two examples of systems of this architecture are
Kepler-48 and Kepler-68. By starting the planets of these sys-
tems in transiting configuration, we force the starting conditions
to be a roughly coplanar disk containing all the planets.
Kepler-48 (Steffen et al. 2013; Marcy et al. 2014) is a four
planet system with three inner transiting planets and one non-
transiting companion at more than 1 AU (a minimum mass 657
M⊕ companion with a period of roughly a 980 day period).
Kepler-68 (Gilliland et al. 2013) is a three planet system with
two transiting planets and one non-transiting planet, also outside
of 1 AU (with a minimum mass of 0.95 Mjup companion in
roughly a 580 day period).
To evaluate the transit stability of Kepler-48 and Kepler-68,
we performed the same Monte Carlo evolution described in Sec-
tion 3.1, with all orbital parameters drawn from observationally
constrained priors except inclination. Though the true orbital
inclination of the outer planets in the Kepler-48 and Kepler-68
systems is not known, we choose the orbital inclinations for the
giant outer planets in each system by drawing a mutual inclina-
tion plane width from a Rayleigh distribution with a width of
1.5◦(from Fabrycky et al. 2014, which suggested a Rayleigh
distribution width between 1.0◦and 2.2◦). We constrain this
choice of plane width such that the planets all start out mutu-
ally transiting, to mimic the starting conditions of the compact
Kepler systems. With these starting conditions, we are probing
what would happen to the observability of these systems over
time, if they did start on feasibly observable architectures.
Through 4000 trials, Kepler-48 and Kepler-68 exhibited
significantly more range in their mutual impact parameters than
the other compact Kepler systems. Figure 6 plots the behavior of
Kepler-48 and Kepler-68 overlaid on the previous result for the
compact Kepler systems. Kepler-68's mean mutual impact pa-
rameter is well above the limit for a mutually transiting system,
while Kepler-48 spends about 60% of its orbits in a transiting
configuration (marginalized over starting parameters).
We treat Kepler-48 and Kepler-68 as isolated systems. In
other words, in our experiments, the only perturbation avail-
able to excite oscillations in inclination is that of the interac-
tions between known bodies in each system. Thus, by generat-
ing the mean mutual impact parameter over one secular period
for these systems after they start in a transiting configuration,
we can make a statement about the amplitude of self-excitation
in these compact systems. As shown in Figure 6, both Kepler-48
and Kepler-68 would be expected to develop significant mutual
inclinations that prevent all planets from being seen in transit
purely through excite self-excitations of inclination. Figure 4
shows as salmon points the mutual transit probability for these
two systems, confirming that it is unlikely that the magnitude
of the secular interactions would allow these two planets to be
seen in transit.
This result indicates that even if these systems were to be-
gin their secular evolution in a roughly coplanar configuration,
they would be expected to self-excite sufficient oscillations to
produce the current orbital state (where not all planets transit
- we do not have sufficient limits on the observed inclinations
Figure 5. For the generalized multi-planet Kepler systems, the param-
eters of the system were sampled 4000 times and evolved forward in
time, just as in Figure 3. The resulting inclination angles for the plan-
etary orbits were converted to a mutual impact parameter (see text).
The mean and scatter of these values are plotted here for each system
as a function of the total mass of the transiting planets, given in earth
masses. The dotted horizontal line indicates the level above which it is
not possible to observe all the planets in transit.
3.3
Inclination Oscillations in Systems with
Non-transiting Planets
Long-term RV followup to systems with transiting planets has
not only found masses for Kepler planets, but has also resulted
in the characterization of additional, non-transiting companions
to some transiting systems (Marcy et al. 2014). Additionally,
transit-timing variation analysis (Agol et al. 2005; Holman &
Murray 2005) has both confirmed masses of planets and pro-
vided additional candidate planets (Cochran et al. 2011; Had-
den & Lithwick 2014). The current state of these systems pro-
vides insight to their dynamical history: assuming that systems
form from roughly coplanar protoplanetary disks, something in
the evolution of these systems has resulted in sufficiently large
spread in inclinations to prevent all planets from being seen in
transit.
As shown in Section 3.1, the observed multi-transiting Ke-
pler systems are CMT-stable against self-perturbation (mutual
inclinations excited by dynamical interactions between the tran-
siting planets). Furthermore, the generalized Kepler systems are
more likely than not to be seen in mutual transit. For multi-
planet systems with some planets transiting and additional non-
transiting companions, something in the dynamical history of
the systems has resulted in misalignment in inclination between
the planets. This effect could be explained in one of many ways:
it could be due to a difference in formation mechanism be-
tween the purely multi-transiting systems and the systems with
some planets outside the transiting plane; it could be due to
some other perturbation, such as an as-yet undiscovered stel-
lar or massive planetary companion (a possibility beyond the
scope of this paper); or finally, it could be due to the effect of
self-excitation between all (known) planets in the system. Our
analysis probes this final possibility, which would apply if all
discovered planets (both those that are currently transiting and
those that are currently non-transiting) in a system had started
out roughly coplanar, in a potentially transiting configuration,
c(cid:13) 2015 RAS, MNRAS 000, 1–14
8
Becker & Adams
Figure 6. Kepler systems in which all discovered planets are transiting
are plotted as black points (they correspond to the same data presented
in Figure 3), while Kepler systems where additional non-transiting plan-
ets have been discovered are plotted as red points. Kepler-48, marked
(Steffen et al. 2013; Marcy et al. 2014) is a four planet system with
three inner transiting planets and one non-transiting companion outside
of 1 AU. Kepler-68 (Gilliland et al. 2013), marked, is a three planet
system with two transiting planets within 0.1 AU and one additional
non-transiting planet at 1.4 AU.
to make a stronger comparison). Kepler-48 and Kepler-68 are
examples of systems that 'make sense' dynamically: it is not
required to add additional effects (such as a perturbing com-
panion or stellar flyby) to their systems to explain their current
non-transiting nature. It is important to note that the outer plan-
ets in these two systems are significantly external to the standard
compact systems described in Section 3.1, which generally fell
within 0.5 AU of their host star. Kepler-48 and -68 have outer
companions at roughly 1.4 and 1.8 AU, respectively. It is pos-
sible that part of the reason for the activity of these systems
is the lower transit probability of these outer companions, but
the presence of Kepler-90 (which has an outer companion semi-
major axis of roughly 1 AU) in the CMT-stable sample indicates
that external companions do not ensure non-transiting configu-
rations.
3.4 Comparison to Numerical Integrations
The discussion thus far has considered inclination oscillations
as described by second-order Laplace-Lagrange theory. Al-
though the amplitudes of the oscillations are small, so that the
second order theory is expected to be accurate, in this section
we compare the results to numerical simulations. These latter
calculations, by definition, include interactions to all orders.
For these compact systems, eccentricity and inclination are
generally low, but to evaluate the error inherent in our second-
order expansion, we evolved each compact system using hybrid
symplectic and Bulirsch-Stoer integrator Mercury6 (Cham-
bers 1999). The numerical integrator should provide the ef-
fectively 'right' answer, and significant deviations between the
second-order theory and numerical results would indicate that
second order secular theory is insufficient to describe the evo-
lution of the orbital architectures. We compared 400 numerical
N-body realizations with 400 secular evolutions (see the visu-
alization of one realization of the comparison in Figure 7) to
Figure 7. An illustrative realization of Kelper-341b, with the result from
the numerical N-body code Mercury6 plotted in black and the secular
theory evolution plotted in red.
compute the deviations plotted in Figure 8, which describe the
mean deviation, in degrees, between secular theory and the nu-
merical results. This comparison yielded a standard deviation
of the difference in inclination angle obtained using secular the-
ory and numerical results; this value was found to be less than
0.01◦.
For our use of second order second theory to be adequate
for further analysis, we would want this variation between the
numerical result and secular result to be much smaller than the
threshold for significant inclination (which can cause a planet
to become non-transiting). The planet in our sample with the
largest semi-major axis and largest number of planets in the sys-
tem, Kepler-11g, orbits a star with a radius R∗ = 0.0053 AU.
This planet would need to attain an inclination of 0.65◦ out of
the plane of the other planets to no longer transit. Planets with
semi-major axes less than this value would need an even larger
range of inclinations to be no longer seen as mutually transiting.
Given that the typical deviation between the Mercury6 nu-
merical results and secular theory is less than 0.01◦, the match
between secular theory and N-body numerics is good enough to
use the second order secular theory for these compact systems.
We additionally note that although there is variation in
the period of secular effects between numerical and second-
order secular theory (Veras & Armitage 2007), this does not
affect our result, as we are concerned with the amplitude rather
than period of inclination oscillations, and these amplitudes are
well-predicted to a reasonable precision. If we were concerned
with the exact period of secular effects, second-order Laplace-
Lagrange theory would not always be sufficient.
Finally, for completeness we note that the standard devia-
tion of the residuals between the secular and numerical results
is not the only measure of the difference (e.g., one could use the
difference between the ranges of inclination angles instead). In
this case, however, the differences between the two approaches
is small: The differences would have to be nearly 100 times
larger in order to change our main conclusion, i.e., that the Ke-
pler compact systems remain CMT-stable.
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Inclination Oscillations for Exoplanet Orbits
9
pled from that of eccentricity, so the null eccentricity approxi-
mation for extracting transit durations from our derived transit
parameters is sufficient. A product of our stability study of the
Kepler systems is time series of I(t) and subsequently ∆b(t).
From these expressions, we can compute the times series τT (t),
evaluated at each transit epoch for each planet in a system.
Thus far, observational study of secular TDVs has been
limited by two main factors: (1) the signature of TDVs caused
by even massive planets is generally small due to small yearly
changes in inclination and eccentricity, and (2) to find TDVs to
good precision, the cadence of photometric measurements must
be high enough such that durations can be extracted from in-
dividual transits. Through TTVs can been used to determine
dynamical quantities of multi-planet systems with good suc-
cess (Agol et al. 2005; Holman & Murray 2005), TDVs in
multi-planet systems are generally as much as several order of
magnitude smaller in amplitude (see, for example, Figure 4 in
Nesvorný et al. 2013, which demonstrates the difference in am-
plitude between a TTV and TDV signal for one system). How-
ever, there has been recent success measuring the amplitude of
planetary TDVs (Eibe et al. 2012). Since transit duration de-
pends on the chord a planet takes across its star in our line of
sight and oscillating inclination can directly change this chord,
secular interactions exciting inclinations will also lead to poten-
tially observable transit duration variations.
Transit duration variations are thought to be one of the few
(but currently feasible) promising ways to find moons around
extrasolar planets (Kipping 2009), as the perturbing effect of a
moon would alter both the time of center transit and the duration
of said transit for a transiting planet. Secular TDVs can also be
used to constrain the oblateness of the central body, which has
been done observationally for the KOI-13 system (Szabó et al.
2012). In this context, the stellar oblateness leads to precession
of the orbital elements and thereby mimics the effects of secu-
lar interactions among multiple planets (see equation 3, which
depends on the stellar oblateness J2). In order for TDVs to be a
useful method to detect exomoons or measure stellar oblateness,
the amplitude due to these effects must be large compared to the
intrinsic variation which we determine here. We also note that
TDVs are now being compiled from the Kepler data (Mazeh et
al. 2013), with more data expected in the near future. The time
series τT (t) yields two useful measures: first, it yields the tran-
sit duration variation rate, which can be parameterized as δτT,t,
the change in duration per unit time (in Table 1, we parame-
terize this as as a variation per year. For example: a TDV of 1
sec yr−1 would mean that over one year, the expected duration
would change by one second, regardless of when or how fre-
quently the transits occur). Second, it yields the duration vari-
ation per orbit, δτT,n, which can be directly compared to the
magnitude of other effects that can also cause TDVs. Both of
these measures provide useful constraints on the properties of
the system: the yearly TDVs provide approximate limits for the
signal due to secular interactions between planets only. The du-
ration variation per orbit allows for a fit to a series of durations
over time, where:
τT (t) = τT (0) + δτT,n n
(18)
where n is the number of orbits observed. If this is done, then
variation accumulates as (∆τT ) = δτT,n n when (∆τT ) is the
total change in duration over an extended baseline of time. In
this case, when the time series contains N independent mea-
surements, the precision in fitting δτT,n, as given in Equation
Figure 8. A residuals plot of (top panel) the deviation in inclination
over several secular periods for each planet in our sample and (bottom
panel) the deviation in eccentricity for the sample sample of realiza-
tions. The averaged deviation in inclination between the numerical and
secular methods is generally below 0.01◦ for all planets.
4 TRANSIT DURATION VARIATIONS
Oscillations of the orbital inclination angles, as described in sec-
ular theory through equation (12), result in planets taking differ-
ent paths across the face of the star as a function of time. These
changing chords, in turn, result in the duration of the planetary
transit varying with inclination and hence with time. For the
case of vanishing eccentricity, we can write τT , the time from
first to fourth contact (the transit duration) for a single transit
analytically (see Seager & Mallén-Ornelas 2003), in the form
τT (t) =
arcsin Θ
P
π
where we have defined the effective angle
(cid:20) (1 + rp/R∗)2 − (a/R∗ cos2 i)
Θ ≡ R∗
a
1 − cos2 i
(cid:21)1/2
(16)
,
(17)
where P is the period of the planet, a is its semi-major axis,
R∗ the radius of the central star, rp is the radius of the planet,
and i is the inclination of the plane; note that the inclination
angle is a function of the time t at which the duration is eval-
uated (so that the duration will also be a function of time). We
also assume that orbital elements are effectively constant dur-
ing a single transit, but that variations occur from transit to tran-
sit. Substituting equation (12) into this expression then yields a
measure of the transit duration, τ, at any point during a planet's
secular evolution. The second order secular theory used in this
work computes motions with the evolution of inclination decou-
c(cid:13) 2015 RAS, MNRAS 000, 1–14
100101102Period,days−0.010−0.0050.0000.0050.0100.0150.0200.0250.0300.035Deviationininclination,degrees(numeric-secular)100101102Period,days−0.00050.00000.00050.00100.00150.0020Deviationineccentricity(numeric-secular)10
Becker & Adams
Figure 9. A histogram of the derived annual TDV values in this work,
for Kepler systems with four or more coplanar planets. Data is presented
in (upper panel) TDV year−1 and (lower panel) TDV orbit−1. The
bulk of the transit duration variations range from 0.01 to 10 seconds per
orbit. The data visualized here is also given in Table 2. This histogram
includes only compact mutually transiting systems with four or more
planets (Kepler-37, -48, and -68 are not included).
(18), is increased. The uncertainty scales like σ ∝ N−3/2, with
one factor of N−1 being due to the number of observed transits,
a factor of N−1/2 being due to the independent nature of these
observations (as used in Pál & Kocsis 2008). In this way, a large
number of transit duration measurements can better constrain
the TDV per orbit than would be possible looking at yearly drift
alone using two widely separated transits (see Figure 3 in Szabó
et al. 2012, which is the first example of observed long-period
TDVs of the type we would see for secular interactions consid-
ered in this work).
The effect of secular interactions between planets in a
multi-planet system can occlude observations of other param-
eters traced by transit durations (such as the presence of exo-
moons or solar oblateness), but it can also provide evidence for
additional planets in the system, as non-transiting planets con-
tribute to the duration variations even if they are not directly
observable.
In Table 2, we present expected yearly TDVs for each
planet considered in this work. These values are also presented
in histograms in Figure 9. Though these values are small be-
cause the yearly change in inclination for each planet is very
small, they provide limits for the kind of TDVs expected in the
observed Kepler multi-planet systems without the presence of
a perturber. The presence of a perturbing secondary in any of
these systems would lead to transit durations outside the ex-
pected range. For example, circumbinary planets can exhibit
TDVs on the order of hours (such as for Kepler-47, as in Welsh
et al. 2014). For exomoons, the TDV amplitude is expected to
−1/2
scale with Msa
, when s denotes a satellite (Kipping 2009).
s
This amplitude is typically on the order of tens of seconds, be-
ing 13.7 seconds for the Earth-Moon system (Kipping 2011). In
comparison, typical values for the secular interactions within
a compact system are a bit smaller (being typically between
10−2 − 101 seconds per orbit).
Significant deviation in transit durations above these pre-
dicted values would suggest the presence of an additional effect
(perturbing planet, extreme solar oblateness, exomoon, etc.) in
the system. The range of transit duration variations summarized
in Figure 9 thus serves as a baseline of the expected TDV dis-
tribution for tightly packed, coplanar, multi-planet systems.
5 PLANETARY MASS CONSTRAINTS
The observed current coplanarity of the Kepler multi-planet sys-
tems is a stringent constraint on the planets' orbital properties.
For most of the planets in the Kepler system, the ratio Rp/R∗
is well-known. Combined with a value of the stellar radius (de-
termined from either spectroscopy or interferometry), this value
yields a measure of the planetary radius.
To perform a dynamical analysis, these measured radii
must be converted to mass. Although some Kepler planets have
masses measured via long-term radial velocity surveys (Marcy
et al. 2014), the population of four-plus planet systems gener-
ally do not have measured masses due to the difficultly of mea-
suring masses for small planets in multi-planet systems. Much
recent work has been conducted aimed at finding a mass-radius
relationship for exoplanets (Wolfgang & Laughlin 2011; Weiss
& Marcy 2014; Rogers 2015). When testing the CMT-stability
of the compact Kepler systems, we use a supplemented version
of the Wolfgang relation (Wolfgang et al. 2015). This relation
introduces a large amount of scatter in density for planets that
could be gaseous or rocky, which is useful for exploring the en-
tire extent of parameter space in which the real planets could
be living. However, another question that the apparent relative
CMT-stability of the Kepler systems engenders is the effect of
systematic mass enhancement (which could be due to an incor-
rect measurement of the stellar radius, as in Muirhead et al.
2012, in which the correction of such a misconception can be
found). To test the effect of such systematic radius errors, we
will inflate the masses of the constituent planets in the Kepler
compact systems and examine the dynamical and CMT-stability
of the systems.
For this experiment, we make a different choice in con-
verting radii
to masses: we use conversion law MP =
M⊕(RP /R⊕)2.1 inferred from results of the Kepler mission
(Lissauer et al. 2011a). Using this relation removes the scatter
due to composition, enabling a qualitative study of the general
stability status of the Kepler multi-planet systems, without noise
from differing compositions between trials.
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Determining the effect of planetary mass enhancement
with respect to roughly estimated values would help determine
if the parameter space of CMT-stable systems (which we have
shown includes all the systems in our sample) changes if the
planetary masses are systematically underestimated. To deter-
mine the extent of this parameter space, we evaluate the dynam-
ical stability of the Kepler systems with varying mass enhance-
ment factors, which places constraints on the maximum ratio by
which the masses can be enhanced without losing the currently
observing transiting configuration of the systems.
To evaluate the effect of having larger planets in each sys-
tem, we performed 40 numerical simulations of each system
using Mercury6 for each mass enhancement factor. The in-
tegration time for each system was 106 dynamical times. This
full treatment accounts for effects ignored in the secular theory
such as the coupling of eccentricity and inclination, and instabil-
ities due to orbit crossing or other effects. A mass enhancement
factor describes the factor by which we increase all planetary
masses within a single system. Although we alter the masses
of the planets, we do not alter starting semi-major axes. The
systems for each enhancement factor were created using ob-
servationally constrained orbital parameters supplemented with
orbital parameters drawn from the standard priors (see Table
3). When a system remains CMT-stable for the entire time, this
means that it is observable in transit and the system as a whole
does not go dynamically unstable (e.g., by ejecting a planet).
There are two potential causes of instability in these sys-
tems. First, increased inclination oscillations can cause a some
planets in a system to lie outside a mutual line of sight, even
as a system remains dynamically stable. For the purposes of
our analysis, we consider this to be an CMT-unstable system.
Second, true dynamical instability (in the form of ejected/star-
consumed planets or orbit crossing) also results in an CMT-
unstable system. When either of these criteria (large inclination
oscillations or true dynamical instability) is met for a certain
mass enhancement factor, we categorize that system as unsta-
ble.
We parameterize the dynamical fullness of a system in
terms of the surface density of a disk consisting of the mass
of its constituent planets spread over an annulus with an inner
radius equal to the semi-major axis of the most interior planet,
and an outer radius equal to the semi-major axis of the most
exterior planet:
(cid:80)i=n
Σ =
i=1 mi
n − a2
1)
π(a2
(19)
where n is the number of planets in a system, a is the semi-
major axis, m is the planetary mass, and i denotes the planet
number.
In Figure 10, we plot the mass enhancement factor required
to make a system CMT-unstable against the the surface density
of the planet annuli. This plot is essentially a comparison of the
dynamical fullness of the system (surface density) to the sta-
bility against excitation (mass enhancement factor required to
knock a system out of transit). The observed result appears to
intuitively support that a higher surface density of material leads
to a less CMT-stable system (for which a lower mass enhance-
ment factor can excited oscillations out of the plane). The large
scatter of the data could also be explained by the existence of
two distinct populations (one containing the disks where planet
surface density is below 200 M⊕/ AU2, and another where the
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Inclination Oscillations for Exoplanet Orbits
11
Figure 10. The mass enhancement factor required to knock a system out
of a CMT-stable transiting configuration, plotted by the surface density
of the annulus containing all the planets (which is defined in Equation
19). The points are shaded based on the ratio of the total planet mass to
stellar mass (Mdisk/M∗). The shape of the trend can be explained two
ways. It could be explained by the existence of two distinct populations
(one containing the disks where planet surface density is below 200
M⊕/ AU2, and another where the density is above 200 M⊕/ AU2,
where the former are significantly less sensitive to mass enhancement),
or it could be explained as a monotonic (but high-scatter) decreasing
trend with surface density.
density is above 200 M⊕/ AU2, where the former are signifi-
cantly less sensitive to mass enhancement).
For many systems with a surface density Σ > 200 M⊕/
AU2, hot or warm Jupiter-like planets would be CMT-stable
even in a multiple-planet system. This finding suggests that
Jovian-size planets can exist in tightly-packed multi-planet sys-
tems with semi-major axis similar to those of the discovered Ke-
pler systems (although this result holds only for Myr timescales,
as discussed below).
The mass enhancement factor required to render the sys-
tems CMT-unstable may seem higher than expected. On one
hand, the integrations are carried out for only 106 dynamical
times, which generally works out to be a few million years,
which is short compared to the system ages. The critical en-
hancement factor appropriate for the ages of the systems are
thus lower, but we assume here that the short-time values pro-
vide a good relative measure of stability. On the other hand,
these systems are in CMT-stable configurations, even though
their surface densities are much larger than that of out solar
system (the analogous value for our solar system is 0.49 M⊕/
AU2). For comparison, we note that the GJ 876 system (one of
the most dynamically active systems discovered to date) has a
surface density Σ = 2750 M⊕/ AU−2, which is much larger
than the systems considered here.
6 CONCLUSIONS
This paper has explored the dynamics of compact solar sys-
tems undergoing oscillations in their orbital inclination angles.
If such oscillations occur with sufficient amplitude, then not all
of the planets in a multi-planet system are expected to tran-
sit at a given epoch. By comparing the conditions required for
the excitation of inclination angles with the observed proper-
ties of compact multi-planet systems, we can put constraints on
0200400600800Surfacedensity,Σ,M⊕AU−2050100150MassEnhancementfactor0.00040.00080.00120.00160.00200.00240.0028Mdisk/M∗12
Becker & Adams
their dynamical history. In this work, we have provided mea-
sures of ∆b(t), the spread in impact parameters, and character-
ized the potential dynamical history of compact extrasolar sys-
tems. We have also utilized our method to test the dynamical
and CMT-stability of a small sample of systems with additional
non-transiting planets. From our derived ∆b(t), we have ex-
tracted subsequently the expected TDVs for observed systems
in the case that these systems have no additional non-transiting
companions. Finally, we have explored the effect of enhancing
the mass of planets in these tightly packed systems, with an aim
at determining how robustly the transit stability holds as plane-
tary masses increase.
We have done this analysis by examining the multi-planet
Kepler systems with the greatest number of transiting planets
and analyzing their long-term stability, using a combination of
secular (Sections 3, 3.3, and 4) and numerical techniques (Sec-
tions 3.4 and 5). Using the Kepler systems with the greatest
number of transiting planets as our sample, we derived ∆b(t)
for each planet using Monte Carlo techniques to marginalize
over potential values of present orbital elements. We have deter-
mined that the compact Kepler systems are CMT-stable against
being excited into non-mutually-transiting configurations.
Compact solar systems could have configurations that al-
low for a significant spread in the orbital inclinations through
secular interactions between the constituent planets (Section 3).
However, for the types of architectures observed in the Kepler
sample of multi-planet systems, the expected range of inclina-
tion angles is almost always small. As shown in Figure 3, the
typical spread in the mean mutual impact parameter is typically
less than ∼ 0.5, whereas impact parameters greater than 2 are
required for planets to move out of transit. This result can also
be expressed in terms of inclination angles: self-excitation gen-
erally produces ∆i <∼ 0.5◦, whereas angles of 1 – 2◦ are re-
quired to compromise transit in these compact systems. As a
result, for most of the systems discovered by the Kepler mis-
sion, the self-excitation of inclination angle oscillations is gen-
erally not large enough to prevent planets from being observed
in transit.
We have also tested the behavior of generalized Kepler sys-
tems. For these generalizations, we drew orbital parameters for
each system from expanded but observationally inspired poste-
riors, then tested the dynamical stability. For dynamically stable
analogs, we proceeded with the analysis used for the observed
Kepler systems. We found that the generalized systems are ex-
perience significantly more action in mutual impact parame-
ter excitation, resulting in these systems being on average less
CMT-stable than the observed Kepler systems. The observed
Kepler systems are remarkably CMT-stable, even compared to
their analogs.
Our derived result that self-excitation of inclination an-
gle oscillations is generally not large enough to prevent planets
from being observed in transit holds for the Kepler systems, but
not their analogs; even then, it has an important exception. We
have also considered another type of Kepler system that con-
tains 2 or 3 transiting planets and an additional planet not seen
in transit (where the additional body was discovered by radial
velocity follow-up). Kepler 48 and Kepler 68 are examples of
this type of system. These systems are CMT-unstable to signifi-
cant oscillations in inclination angle, so that the expected spread
in inclination angle is generally large enough to move planets
out of transit. We found this result by secularly evolving these
systems after starting them in a nearly coplanar configuration.
Even starting roughly coplanar, the magnitude of these systems'
self-excitation is large enough that not all planets can be seen in
transit simultaneously for most of each system's orbital history.
This finding indicates that the current Kepler systems with non-
transiting companions could have started roughly coplanar and
subsequently had some of their planets excited out of the plane
via dynamical interactions between the planets that we know
about. Specifically, it is not necessary to introduce additional
bodies into these systems to recreate the currently observed ar-
chitectures.
We have focused on the secular interactions of compact
systems of planets, and derived observables corresponding to
the current known properties of these systems. These observ-
ables, the transit duration variations for Kepler systems with the
observationally determined properties, are given in Table 2. Im-
plicit in the motivation behind the calculation of these TDVs
is the idea that there could be additional bodies in the systems
we are considering, leading to true TDVs deviating from those
that we have found here. An additional massive companion or
an exomoon, for example, could cause transit duration varia-
tions with a larger amplitude than those derived in this work. If
future observations of TDVs in these systems are vastly differ-
ent than expected, it could potentially be evidence for either an
exomoon or additional, exterior, non-transiting bodies in these
compact systems.
We have also explored the effect of planetary mass en-
hancement in these systems. The stability of systems is related
to how much the constituent planets' masses must be enhanced
to result in a system that will no longer mutually transit. Gen-
erally, systems with higher effective surface density (calculated
by spreading the mass of discovered planets within an annulus
with inner and outer radii equal to the inner and outer planet's
orbital radii) do not allow mass enhancement factors as high as
those with lower surface density. This result suggests that dy-
namically 'full' systems would not be mutually transiting if they
hosted Jovian-mass planets. However, some systems with lower
surface densities would be CMT-stable in a transiting configura-
tion even with Jovian-mass planets (at least over time scales of
∼ 10 Myr), indicating that it might be possible to see multi-
transiting compact systems with Jovian-mass planets if they
existed. The stability boundaries – over longer time scales –
should be explored further in future work.
Spreads in the inclination angles in compact systems can
be produced by a variety of astronomical processes, in addition
to those considered in this work. Excitation by the compact so-
lar system planets themselves (with semi-major axes a <∼ 0.5
AU) is not generally a significant effect, but we have not (yet)
calculated the effect caused by possible additional bodies in the
outer part of the solar system (where a ≈ 5 − 30 AU). Since
planet formation is a relatively efficient process, the additional
giant planets, not seen in transit by Kepler, are not only possi-
ble but likely. The orbits of these outer planets can be endowed
with high inclination angles through a variety of dynamical
mechanisms. For example, most solar systems form within clus-
ters, and inclinations can be excited through dynamical interac-
tions between solar systems and other cluster members (Adams
& Laughlin 2001; Malmberg et al. 2007; Adams 2010; Li &
Adams 2015). In addition, a range of inclination angles can be
realized through the formation of planets in warped disks. The
observed angular momentum vectors in star-forming cores do
not point in the same direction as a function of radius (Good-
man et al. 1993; Caselli et al. 2002). This heterogeneity can lead
c(cid:13) 2015 RAS, MNRAS 000, 1–14
to differences in angular momentum vector of the disk plane
as a function of radius (for disks produced through collapse of
the cores), which in turn will influence the inclination angles of
forming planets (see also Spalding et al. 2014). These various
mechanisms can lead to inclined, massive, outer secondaries to
the compact systems that we have considered in this work. The
presence of such secondaries would alter the stability of these
systems, and this effect could be evident in the TDVs. Addi-
tionally, it is possible that a system of planets would have only
some planets mutually transiting, instead of the condition of all
planets in a system transiting that we have considered in this
work. For example, although we see four planets in a system dis-
covered by Kepler, it is possible that another short-period com-
panions exists in such a system, resulting in our picture of the
system being incomplete. Extensions on our calculation that ac-
count for this possibility could potentially be explored by using
techniques such as the semi-analytical code CORBITS (Brak-
ensiek & Ragozzine, 2015).
In summary, we have determined that self-excitation is not
usually a dominant mechanism in exciting mutual inclination in
tightly packed, multi-planet systems. Self-excitation does op-
erate in some solar system architectures, where Kepler-48 and
Kepler-68 are prime examples. Subsequent analysis of the effect
of perturbing secondaries and stellar fly-bys in a dense cluster
environment will complete the picture of how and when mutual
inclinations are excited in exoplanetary systems.
ACKNOWLEDGEMENTS
We would like to thank Konstantin Batygin, Kathryn Volk, Ben
Montet, Andrew Vanderburg, and Doug Lin for useful conver-
sations. We would like to additionally thank Konstantin Baty-
gin for his careful review of the manuscript and helpful sug-
gestions. We would like to thank the referee, Darin Ragozzine,
for his thoughtful and helpful suggestions. J.B. is supported by
the National Science Foundation Graduate Research Fellow-
ship, Grant No. DGE 1256260.
REFERENCES
Adams, F. C. 2010, ARA&A, 48, 47
Adams, F. C., & Laughlin, G. 2001, Icarus, 150, 151
Agol, E., Steffen, J., Sari, R., & Clarkson, W. 2005, MNRAS,
359, 567
Ballard, S., & Johnson, J. A. 2014, arXiv:1410.4192
Barclay, T., Rowe, J. F., Lissauer, J. J., et al. 2013, Nature, 494,
452
Barclay, T., Quintana, E. V., Adams, F. C., et al. 2015,
arXiv:1505.01845
Batalha, N. M., Rowe, J. F., Bryson, S. T., et al. 2013, ApJS,
204, 24
Batygin, K., & Adams, F. C. 2013, ApJ, 778, 169
Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 736,
19
Borucki, W. J., Agol, E., Fressin, F., et al. 2013, Science, 340,
587
Brakensiek, J., & Ragozzine, D. 2015, submitted
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Inclination Oscillations for Exoplanet Orbits
13
Caselli, P., Benson, P. J., Myers, P. C., & Tafalla, M. 2002,
ApJ, 572, 238
Chambers, J. E. 1999, MNRAS, 304, 793
Chiang, E., & Laughlin, G. 2013, MNRAS, 431, 3444
Cochran, W. D., Fabrycky, D. C., Torres, G., et al. 2011, ApJS,
197, 7
Dawson, R. I., Johnson, J. A., Morton, T. D. et al. 2012, ApJ,
761, 163
Dermott, S. F. 1973, Nature Physical Science, 344, 18
Eibe, M. T., Cuesta, L., Ullán, A., Pérez-Verde, A., & Navas,
J. 2012, MNRAS, 423, 1381
Fabrycky, D. C., & Winn, J. N. 2009, ApJ, 696, 1230
Fabrycky, D. C., Ford, E. B., Steffen, J. H. et al. 2012, ApJ,
750, 114
Fabrycky, D. C., Lissauer, J. J., Ragozzine, D., et al. 2014, ApJ,
790, 146
Fang, J., & Margot, J.-L. 2012, ApJ, 761, 92
Fielding, D. B., McKee, C. F., Socrates, A., Cunning-
ham, A. J., & Klein, R. I. 2014, submitted to MNRAS,
arXiv:1409.5148
Gilliland, R. L., Marcy, G. W., Rowe, J. F., et al. 2013, ApJ,
766, 40
Goodman, A. A., Barranco, J. A., Fuller, G. A., & Myers, P. C.
1993, ApJ, 406, 528
Hadden, S., & Lithwick, Y. 2014, ApJ, 787, 80
Holman, M. J., & Murray, N. W. 2005, Science, 307, 1288
Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126,
398
Jensen, E. L., & Akeson, R. 2014, Nature, 511, 567
Kipping, D. M. 2009, MNRAS, 392, 181
Kipping, D. M. 2011, The Transits of Extrasolar Planets
with Moons, by Kipping, David M. ISBN: 978-3-642-22268-
9. Berlin: Springer, 2011,
Kozai, Y. 1962, AJ, 67, 591
Lada, C. J., & Lada, E. A. 2003, ARA&A, 41, 57
Laughlin, G., Crismani, M., & Adams, F. C. 2011, ApJ, 729,
L7
Li, G., & Adams, F. C. 2015, MNRAS, 448, 344
Lidov, M. L. 1962, Planet. Space Sci., 9, 719
Lissauer, J. J., Ragozzine, D., Fabrycky, D. C. et al. 2011,
ApJS, 197, 8
Lissauer, J. J., Fabrycky, D. C., Ford, E. B., et al. 2011, Nature,
470, 53
Lovis, C., Ségransan, Mayor, M. et al. 2011, A&A, 528, 112
Malmberg, D., de Angeli, F., Davies, M. B., Church, R. P.,
Mackey, D., & Wilkinson, M. I. 2007, MNRAS, 378, 1207
Marcy, G. W., Isaacson, H., Howard, A. W., et al. 2014, ApJs,
210, 20
Mazeh, T., Nachmani, G., Holczer, T., et al. 2013, ApJS, 208,
16
Morton, T. D., & Winn, J. N. 2014, ApJ, 796, 47
Muirhead, P. S., Johnson, J. A., Apps, K., et al. 2012, ApJ, 747,
144
Murray, C. D., & Dermott, S. F. 1999, Solar System Dynamics
(Cambridge: Cambridge Univ. Press)
Nesvorný, D., Kipping, D., Terrell, D., et al. 2013, ApJ, 777, 3
Pál, A., & Kocsis, B. 2008, MNRAS, 389, 191
Pfalzner, S. 2013, A&A, 549, 82 Early evolution of the birth
cluster of the solar system
Porras, A., et al. 2003, AJ, 126, 1916
Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2014, Proc.
SPIE, 9143, 914320
14
Becker & Adams
Ragozzine, D., & Holman, M. J. 2010, arXiv:1006.3727
Rogers, L. A. 2015, ApJ, 801, 41
Rowe, J. F., Bryson, S. T., Marcy, G. W., et al. 2014, ApJ, 784,
45
Seager, S., & Mallén-Ornelas, G. 2003, ApJ, 585, 1038
Spalding, C., Batygin, K., & Adams, F. C. 2014, ApJ, 797,
L29, arXiv:1411.5431.
Szabó, G. M., Pál, A., Derekas, A., et al. 2012, MNRAS, 421,
L122
Steffen, J. H., Fabrycky, D. C., Agol, E., et al. 2013, MNRAS,
428, 1077
Van Laerhoven, C., & Greenberg, R. 2012, Celestial Mechan-
ics and Dynamical Astronomy, 113, 215
Veras, D., & Armitage, P. J. 2007, ApJ, 661, 1311
Walkowicz, L. M., & Basri, G. S. 2013, MNRAS, 436, 1883
Weiss, L. M., & Marcy, G. W. 2014, ApJ, 783, LL6
Welsh, W. F., Orosz, J. A., Carter, J. A., & Fabrycky, D. C.
2014, IAU Symposium, 293, 125
Wolfgang, A., & Laughlin, G. 2011, arXiv:1108.5842
Wolfgang, A., Rogers, L. A., & Ford, E. B. 2015,
arXiv:1504.07557
Wu, Y., & Lithwick, Y. 2013, ApJ, 772, 74
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Inclination Oscillations for Exoplanet Orbits
15
Transit Duration Variations for Kepler Compact Systems
Planet
Orbital Period, days
τT,n, TDV (s yr−1)
τT,t, TDV (s orbit−1)
Kepler 11
Kepler 11
Kepler 11
Kepler 11
Kepler 11
Kepler 11
Kepler 20
Kepler 20
Kepler 20
Kepler 20
Kepler 20
Kepler 24
Kepler 24
Kepler 24
Kepler 24
Kepler 26
Kepler 26
Kepler 26
Kepler 26
Kepler 32
Kepler 32
Kepler 32
Kepler 32
Kepler 32
Kepler 33
Kepler 33
Kepler 33
Kepler 33
Kepler 33
Kepler 49
Kepler 49
Kepler 49
Kepler 49
Kepler 55
Kepler 55
Kepler 55
Kepler 55
Kepler 55
Kepler 62
Kepler 62
Kepler 62
Kepler 62
Kepler 62
Kepler 79
Kepler 79
Kepler 79
Kepler 79
Kepler 80
Kepler 80
Kepler 80
Kepler 80
Kepler 82
Kepler 82
Kepler 82
Kepler 82
Kepler 84
Kepler 84
Kepler 84
Kepler 84
Kepler 84
Kepler 85
Kepler 85
Kepler 85
Kepler 85
10.3039
13.0241
22.6845
31.9996
46.6888
118.3807
3.6961219
6.098493
10.854092
19.57706
77.61184
4.244384
8.1453
12.3335
18.998355
3.543919
12.2829
17.2513
46.827915
0.74296
2.896
5.90124
8.7522
22.7802
5.66793
13.17562
21.77596
31.7844
41.02902
2.576549
7.2037945
10.9129343
18.596108
2.211099
4.617534
10.198545
27.9481449
42.1516418
5.714932
12.4417
18.16406
122.3874
267.291
13.4845
27.4029
52.0902
81.0659
3.072186
4.645387
7.053
9.522
2.382961
5.902206
26.444
51.538
4.224537
8.726
12.883
27.434389
44.552169
8.306
12.513
17.91323
25.216751
c(cid:13) 2015 RAS, MNRAS 000, 1–14
9.322660729
8.478858911
27.49182915
9.333648315
47.07565441
195.5615046
1.914759018
9.46333942
4.0055073
50.15850045
73.36444964
0.743325382
1.056218584
1.232153241
3.605506001
1.859049714
3.710922038
6.100397395
37.7182322
2.028539272
0.980507587
1.7046195
3.609925168
5.932320689
4.848962172
5.861904438
2.870395649
10.62434609
10.88222417
1.222377872
2.229154696
3.964566165
16.67313555
0.74535018
2.075484138
7.323822446
7.955549215
23.59027592
6.558019889
10.787751
7.847335247
126.4774292
200.7871286
12.8033798
28.46032201
20.63076067
89.08814913
2.130009061
1.943170254
1.714628021
1.971262692
0.965129741
2.222408585
17.76864363
10.89753388
4.439153917
2.002291695
3.658674592
10.50305215
29.82330904
1.592408413
2.059758945
15.62065426
26.0037509
0.263177435
0.302546593
1.708598352
0.818282226
6.021659764
63.42659673
0.019389542
0.158115368
0.119112725
2.690290337
15.59986281
0.008643722
0.023570458
0.041634964
0.187667625
0.018050196
0.124879135
0.28832818
4.839085401
0.004129106
0.007779589
0.027559914
0.08656106
0.370245073
0.075297474
0.211600617
0.171248276
0.925173879
1.223252036
0.008628812
0.043995541
0.118534384
0.84946693
0.004515186
0.026256489
0.204636528
0.609158472
2.724298248
0.102681199
0.367720443
0.390519091
42.40888689
147.0372394
0.473005959
2.136699611
2.944275204
19.78633147
0.017928175
0.024730898
0.03313225
0.051425653
0.006301004
0.035937297
1.287326061
1.538731784
0.051379096
0.047868486
0.129136177
0.789437858
3.640255081
0.036237108
0.070613051
0.766620199
1.796520854
16
Becker & Adams
Transit Duration Variations for Kepler Compact Systems (continued)
Planet
Orbital Period, days
τT,n, TDV (s yr−1)
τT,t, TDV (s orbit−1)
Kepler 90
Kepler 90
Kepler 90
Kepler 90
Kepler 90
Kepler 90
Kepler 90
Kepler 102
Kepler 102
Kepler 102
Kepler 102
Kepler 102
Kepler 106
Kepler 106
Kepler 106
Kepler 106
Kepler 107
Kepler 107
Kepler 107
Kepler 107
Kepler 122
Kepler 122
Kepler 122
Kepler 122
Kepler 150
Kepler 150
Kepler 150
Kepler 150
Kepler 169
Kepler 169
Kepler 169
Kepler 169
Kepler 169
Kepler 172
Kepler 172
Kepler 172
Kepler 172
Kepler 186
Kepler 186
Kepler 186
Kepler 186
Kepler 186
Kepler 197
Kepler 197
Kepler 197
Kepler 197
Kepler 208
Kepler 208
Kepler 208
Kepler 208
Kepler 215
Kepler 215
Kepler 215
Kepler 215
Kepler 220
Kepler 220
Kepler 220
Kepler 220
Kepler 221
Kepler 221
Kepler 221
Kepler 221
7.008151
8.719375
59.73667
91.93913
124.9144
210.60697
331.60059
5.28696
7.07142
10.3117
16.1457
27.4536
6.16486
13.5708
23.9802
43.8445
3.179997
4.901425
7.958203
14.749049
5.766193
12.465988
21.587475
37.993273
3.428054
7.381998
12.56093
30.826557
3.250619
6.195469
8.348125
13.767102
87.090195
2.940309
6.388996
14.627119
35.118736
3.8867907
7.267302
13.342996
22.407704
129.9441
5.599308
10.349695
15.677563
25.209715
4.22864
7.466623
11.131786
16.259458
9.360672
14.667108
30.864423
68.16101
4.159807
9.034199
28.122397
45.902733
2.795906
5.690586
10.04156
18.369917
44.64259007
51.72347308
122.8193078
163.2018666
196.2255593
160.6239691
200.873026
21.67804261
5.624323905
3.349150204
0.736673755
42.68179251
3.817466082
7.56602341
35.33153826
18.90988445
0.746863983
0.979622681
4.806241849
1.19795885
3.958098992
0.796688261
24.6377173
64.29884982
1.867968128
1.135021003
4.078868983
18.32964457
2.003755495
3.473185364
4.542706201
4.044452569
21.57958002
0.46849388
0.965003916
3.850997739
8.863030127
2.174369234
2.671048208
5.6398171
21.23861165
127.2638103
1.746693303
1.664512376
2.946426313
19.27829892
0.327987695
1.085357765
2.145926971
1.939023031
5.214125094
7.140540403
13.54786835
73.36673268
1.08595549
0.726348102
31.65200923
28.54622634
1.622302263
1.258209005
3.116385606
7.160429778
0.857156198
1.235606461
20.10086701
41.10859624
67.15451508
92.6808971
182.492093
0.314002587
0.108964265
0.094617622
0.032586612
3.210325641
0.064477107
0.281306823
2.321253024
2.271491586
0.006506918
0.013154924
0.104791913
0.048407545
0.062529213
0.027209606
1.457167415
6.692941794
0.017543824
0.022955405
0.140368186
1.548054337
0.017845057
0.058953458
0.103898847
0.152549017
5.148958444
0.003774019
0.016891524
0.154326033
0.852762781
0.023154296
0.053181682
0.206170019
1.303858968
45.30734602
0.026795271
0.047197796
0.126555573
1.331508004
0.003799841
0.022202623
0.065446575
0.086376612
0.133719767
0.286934458
1.145608602
13.70068658
0.012376343
0.017978009
2.438713341
3.589999468
0.012426862
0.019616292
0.085735269
0.360373975
c(cid:13) 2015 RAS, MNRAS 000, 1–14
Transit Duration Variations for Kepler Compact Systems (continued)
Planet
Orbital Period, days
τT,n, TDV (s yr−1)
τT,t, TDV (s orbit−1)
Inclination Oscillations for Exoplanet Orbits
17
Kepler 223
Kepler 223
Kepler 223
Kepler 223
Kepler 224
Kepler 224
Kepler 224
Kepler 224
Kepler 235
Kepler 235
Kepler 235
Kepler 235
Kepler 238
Kepler 238
Kepler 238
Kepler 238
Kepler 238
Kepler 251
Kepler 251
Kepler 251
Kepler 251
Kepler 256
Kepler 256
Kepler 256
Kepler 256
Kepler 265
Kepler 265
Kepler 265
Kepler 265
Kepler 282
Kepler 282
Kepler 282
Kepler 282
Kepler 286
Kepler 286
Kepler 286
Kepler 286
Kepler 296
Kepler 296
Kepler 296
Kepler 296
Kepler 296
Kepler 299
Kepler 299
Kepler 299
Kepler 299
Kepler 306
Kepler 306
Kepler 306
Kepler 306
Kepler 338
Kepler 338
Kepler 338
Kepler 338
Kepler 341
Kepler 341
Kepler 341
Kepler 341
Kepler 402
Kepler 402
Kepler 402
Kepler 402
7.384108
9.848183
14.788759
19.721734
3.132924
5.925003
11.349393
18.643577
3.340222
7.824904
20.060548
46.183669
2.090876
6.155557
13.233549
23.654
50.447
4.790936
16.514043
30.133001
99.640161
1.620493
3.38802
5.839172
10.681572
6.846262
17.028937
43.130617
67.831024
9.220524
13.638723
24.806
44.347
1.796302
3.468095
5.914323
29.221289
5.841648
10.21457
19.850242
34.14204
63.336
2.927128
6.885875
15.054786
38.285489
4.646186
7.240193
17.326644
44.840975
9.341
13.726976
24.310856
44.431014
5.195528
8.01041
27.666313
42.473269
4.028751
6.124821
8.921099
11.242861
c(cid:13) 2015 RAS, MNRAS 000, 1–14
6.794688993
4.741440907
3.494873231
11.9376658
1.856809099
0.971515147
3.646370338
13.12338292
0.564712568
7.345399102
6.816955222
39.50423791
0.884749492
2.172310058
3.287882421
1.177061595
39.55458755
3.555780556
6.199763248
10.65335888
96.79454965
0.323635851
0.427493436
0.884567462
2.074977703
3.529407786
6.886638378
31.42320459
60.40531736
15.78120361
20.28678008
25.75148128
26.5277244
0.713292306
1.26801456
3.526176454
13.69225381
2.228023946
15.68127047
16.85583517
64.98518257
82.07825957
1.224694325
1.810056081
12.26546764
27.25944274
3.606915153
3.698312183
10.59158411
39.39149661
3.640821013
3.068493345
7.040320215
13.81305425
1.217989712
1.039299363
22.75719988
12.26030004
1.081566673
0.917643194
3.453650264
2.886419391
0.1374595
0.12793035
0.141602296
0.645017725
0.015937649
0.015770494
0.113381068
0.67032
0.00516785
0.15747135
0.374662623
4.998494925
0.005068223
0.036635009
0.119206447
0.076280041
5.466877474
0.04667265
0.2805018
0.879500476
26.4236288
0.001436848
0.003968099
0.014151073
0.060723353
0.066200686
0.32129351
3.713156719
11.22562885
0.398660183
0.758043217
1.750112999
3.223082175
0.003510379
0.012048205
0.05713684
1.09617892
0.035658443
0.438842287
0.916691527
6.078703295
14.24248945
0.009821471
0.034147452
0.505901344
2.859290672
0.045913421
0.073360257
0.502785225
4.839323602
0.093175093
0.115400369
0.468921126
1.681446594
0.017337259
0.022808806
1.724952917
1.426671292
0.01193798
0.015398357
0.084411934
0.088908526
18
Becker & Adams
Transit Duration Variations for Kepler Compact Systems (continued)
Planet
Orbital Period, days
τT,n, TDV (s yr−1)
τT,t, TDV (s orbit−1)
Kepler 444
Kepler 444
Kepler 444
Kepler 444
Kepler 444
3.6001053
4.5458841
6.189392
7.743493
9.740486
2.256350153
1.40080575
1.916669517
2.784776198
1.368811474
0.022255063
0.017446303
0.032501422
0.059079164
0.036528463
Table 2. Predicted values of the transit duration variations (TDVs) for the current sample of Kepler compact systems containing only the planets
that have been discovered so far. Duration variations are presented both per orbit as well as per year. Errors are typically on the order of 1% of
reported values, but are not reported for brevity.
c(cid:13) 2015 RAS, MNRAS 000, 1–14
|
1305.3795 | 1 | 1305 | 2013-05-16T13:27:36 | Lucky Imaging of transiting planet host stars with LuckyCam | [
"astro-ph.EP",
"astro-ph.SR"
] | We obtained high-resolution, high-contrast optical imaging in the SDSS $i'$ band with the LuckyCam camera mounted on the 2.56m Nordic Optical Telescope, to search for faint stellar companions to 16 stars harbouring transiting exoplanets. The Lucky Imaging technique uses very short exposures to obtain near diffraction-limited images yielding sub-arcsecond sensitivity, allowing us to search for faint stellar companions within the seeing disc of the primary planet host. Here we report the detection of two candidate stellar companions to the planet host TrES-1 at separations $<6.5\arcsec$ and we confirm stellar companions to CoRoT-2, CoRoT-3, TrES-2, TrES-4, and HAT-P-7 already known in the literature. We do not confirm the candidate companions to HAT-P-8 found via Lucky Imaging by \citet{Bergfors2013}, however, most probably because HAT-P-8 was observed in poor seeing conditions. Our detection sensitivity limits allow us to place constraints on the spectral types and masses of the putative bound companions to the planet host stars in our sample. If bound, the stellar companions identified in this work would provide stringent observational constraints to models of planet formation and evolution. In addition these companions could affect the derived physical properties of the exoplanets in these systems. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 27 July 2021
(MN LATEX style file v2.2)
Lucky Imaging of transiting planet host stars with
LuckyCam
F. Faedi1(cid:63), T. Staley2, Y. G´omez Maqueo Chew3,1, D. Pollacco1,
S. Dhital3, S. C. C. Barros5, I. Skillen6, L. Hebb3, C. Mackay2, and C. A. Watson7
1Department of Physics, University of Warwick, Coventry CV4 7AL, UK
2Institute of Astronomy, Cambridge University, Madingley Road, Cambridge CB3 0HA, UK
3Department of Physics and Astronomy, Vanderbilt University, Nashville, TN 37235, USA
4Department of Astronomy, Boston University, 725 Commonwealth Avenue, Boston, MA 02215, USA
5Aix Marseille Universit´e, CNRS, LAM (Laboratoire d'Astrophysique de Marseille) UMR 7326, 13388, Marseille, France
6Isaac Newton Group of Telescope, Apartado de Correos 321, Santa Cruz de la Palma, 38700, Spain
7Astrophysics Research Centre, Queen's University Belfast, University Road, Belfast BT7 1NN, UK
ABSTRACT
We obtained high-resolution, high-contrast optical imaging in the SDSS i(cid:48) band with
the LuckyCam camera mounted on the 2.56m Nordic Optical Telescope, to search
for faint stellar companions to 16 stars harbouring transiting exoplanets. The Lucky
Imaging technique uses very short exposures to obtain near diffraction-limited images
yielding sub-arcsecond sensitivity, allowing us to search for faint stellar companions
within the seeing disc of the primary planet host. Here we report the detection of two
candidate stellar companions to the planet host TrES-1 at separations < 6.5(cid:48)(cid:48) and
we confirm stellar companions to CoRoT-2, CoRoT-3, TrES-2, TrES-4, and HAT-P-
7 already known in the literature. We do not confirm the candidate companions to
HAT-P-8 found via Lucky Imaging by Bergfors et al. (2013), however, most probably
because HAT-P-8 was observed in poor seeing conditions. Our detection sensitivity
limits allow us to place constraints on the spectral types and masses of the putative
bound companions to the planet host stars in our sample. If bound, the stellar com-
panions identified in this work would provide stringent observational constraints to
models of planet formation and evolution. In addition these companions could affect
the derived physical properties of the exoplanets in these systems.
Key words:
stars: binaries -- stars: planetary systems
instrumentation: high angular resolution -- methods: observational --
1 INTRODUCTION
More than 800 extrasolar planets have been discovered to
date showing a large variety of physical and dynamical prop-
erties that are dramatically different from those observed in
our Solar System. This has revolutionised our understanding
of planetary formation, structure and evolution. One third of
the known gas giant planets orbit their host at separations
smaller than a few tenths of an AU (with orbital periods
P < 10 d). Among these, transiting systems are specially
important as they allow accurate measurements of masses,
radii, and hence densities, to be derived. These key param-
eters inform us of the system's physical properties, and can
constrain theoretical evolutionary models (e.g. Guillot 2005;
Fortney et al. 2007; Burrows et al. 2007; Liu et al. 2008). In
(cid:63) E-mail: [email protected] -- Part of this work was carried
out while at Queens University Belfast 7
c(cid:13) 0000 RAS
contrast to the planets in our Solar system, exoplanets show
a large variety of orbital properties, for example their orbital
eccentricities span a wide range e = 0 -- 0.97 (e.g. HD 80606,
Pont et al. 2009 and Eggenberger et al. 2004; HAT-P-13,
Bakos et al. 2009). The close-in "hot Jupiters" show a large
angular distribution of (mis)alignments with respect to their
host stars' rotation axis (Winn et al. 2010, 2011, Triaud et al.
2010, Morton & Johnson 2011), and some exoplanets even
have retrograde orbits (e.g. WASP-17, Anderson et al. 2010).
To explain the observed exoplanet orbital configura-
tions, different scenarios have been proposed for migrating
the planets inward from beyond the snow line to their ob-
served position. These migration mechanisms make different
predictions about the current orbital configurations of the
planetary systems. For example, planet-disc interaction via
angular momentum exchange (e.g. Lin et al. 1996, and Ida
& Lin 2004) results in damping any initial inclination of the
planetary orbit with respect to the disc (see e.g., Marzari
2
F. Faedi et al.
& Nelson 2009; Watson et al. 2011). Alternatively, gravi-
tational interaction among multiple giant planets (planet-
planet scattering; e.g. Wu & Murray 2003, Nagasawa et al.
2008), and perturbations induced by a companion star or
a more distant massive planet (Kozai mechanisms, see Fab-
rycky & Tremaine 2007) result in orbital configurations with
large spin-orbit misalignments and large eccentricities. (Ra-
sio & Ford 1996; Weidenschilling & Marzari 1996, Chatterjee
et al. 2008).
Observational evidence for planet-disc migration is
found in multi-planetary systems with mean-motion reso-
nant orbits (e.g. GJ 876, Lee & Peale 2002; Crida et al.
2008). On the other hand, measurements of the Rossiter-
Mclaughlin effect1 (Rossiter 1924; McLaughlin 1924) suggest
that ∼40% of transiting planets have highly tilted orbits pro-
viding supporting evidence for planet-planet scattering and
the Kozai migration mechanism (Winn et al. 2009; Winn
2010). Examples of systems with large spin-orbit misalign-
ments and/or high eccentricities are respectively WASP-17b,
(Anderson et al. 2010), HAT-P-7b (e.g., Winn et al. 2009),
and HD80806b (e.g., Pont et al. 2009; Eggenberger et al.
2004; H´ebrard et al. 2010). More recently Albrecht et al.
(2012) suggested that the Kozai mechanism is responsible
for the migration of the majority, if not all, hot Jupiters,
both mis-aligned and aligned, and that star-planet tidal in-
teraction plays a central role in shaping exoplanets orbital
configurations.
In this paper we present high-contrast, high-angular
resolution optical imaging for 16 stars harbouring transit-
ing extrasolar planets to search for faint stellar compan-
ions. Identifying binary companions to known planet hosts
can provide observational evidence to constrain the different
formation and evolution scenarios, as well as provide crucial
information for subsequent exoplanet characterisation (see
also Bergfors et al. 2013; Daemgen et al. 2009; Narita et al.
2012). The presence of a close-in stellar source to a transiting
planet host star, as in the case for WASP-12 (Bergfors et al.
2013, via Lucky Imaging), could affect the derived planetary
parameters by diluting the transit signal (see also Daem-
gen et al. 2009). For example, Crossfield et al. (2012) find
that WASP-12b is rather hotter and slightly larger (by 1 --
2%) than previously reported, highlighting the importance
of high-resolution imaging for the characterisation of known
and newly discovered transiting planetary systems. Addi-
tionally, the presence of an M dwarf only 1(cid:48)(cid:48) from WASP-12
might have contaminated past atmospheric measurements,
possibly challenging the detection of a high atmospheric
C/O ratio for WASP-12b (Madhusudhan et al. 2011, and
Crossfield et al. 2012, for a recent re-analysis).
The paper is organised as follows: in §2 we briefly de-
scribe our Lucky Imaging technique; §3 presents our Luck-
yCam observations; in §4 we explain the data reduction,
image analysis and candidate detection. Our results are pre-
sented in §5, including the non-detections in our sample. In
§6, we discuss the likelihood of the detected companions be-
ing bound to the planet hosts. Finally, we summarise our
findings and conclusions in §7.
1 See
Encyclopaedia;
http://www.aip.de/People/rheller/content/main spinorbit.html
Holt -- Rossiter -- McLaughlin
the
2 LUCKY IMAGING TECHNIQUE
Lucky Imaging consists of the acquisition of short expo-
sures, at a rate of a few tens of frames per second, using
a very low-noise electron multiplying CCD camera (Fried
1978; Baldwin et al. 2001; Tubbs et al. 2002; Mackay et al.
2004; Law et al. 2006). This allows the rapid image mo-
tion due to atmospheric turbulence to be corrected. Because
the perturbations introduced by the atmosphere change on
timescales of a few milliseconds (known as the atmospheric
coherence time), with fast imaging each frame captures a
different point spread function (PSF) resulting from the at-
mospheric turbulence at that particular moment. By moni-
toring the rapid PSF variations, we can select high quality
short exposures from moments of excellent seeing. During
data reduction the best frames are selected, aligned and co-
added to produce a final image with a bright diffraction
limited core surrounded by a fainter seeing halo. Law et al.
(2006) give a detailed explanation on the Lucky Imaging
technique and the LuckyCam specifications.
3 OBSERVATIONS
Observations were obtained between July 18 and July 22,
2009 at the 2.56m Nordic Optical Telescope (NOT) at
the Roque de los Muchachos Observatory, La Palma, with
the Cambridge LuckyCam vistor instrument. Seeing ranged
from ∼ 0.6(cid:48)(cid:48) to ∼ 1.65(cid:48)(cid:48) as measured by the DIMM (at
500nm; Tokovinin 2002). All observations were made in the
SDSS i(cid:48) band, using a plate scale of 32.4 mas/pixel, provid-
ing good sampling of the PSF. The camera frame rate was
20.75 frames per second using full chip readout (1024 pixels
squared). Table 1 presents a summary of our observations.
Targets were often observed slightly off-centre on the CCD
detector to achieve better positioning of the mosaic field of
view for astrometric calibration. The target observed closest
to the CCD edge has an unbroken observation area of radius
6.5(cid:48)(cid:48). Therefore, to give a uniform dataset we only list de-
tections within 6.5(cid:48)(cid:48). However, we note that the planet host
HAT-P-1 (Bakos et al. 2007) is known to be part of a binary
system with a companion at ∼11(cid:48)(cid:48), that was clearly detected
in our images at a separation r = 11.26 ± 0.03(cid:48)(cid:48), although
this target is not discussed further in this paper.
We selected our sample to optimise the number of
planet host stars observable as by July 2009, in order to
cover a large parameter space of different stellar and plane-
tary properties. Detailed information on individual objects
is available from the Exoplanet Encyclopaedia 2. We present
our sample in Table 1, separating the planet host stars with
candidate companions detected in this work from those with-
out detections.
4 DATA REDUCTION, IMAGE ANALYSIS
AND CANDIDATE DETECTION
4.1 Data Reduction
The data were reduced using the LuckyCam pipeline. Stan-
dard bias correction, gain calibration, and cosmic ray re-
2 http://exoplanet.eu/
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Lucky Imaging of transiting planet hosts
3
Figure 1. Images of TrES-2 obtained with LuckyCam showing the image quality improvement due to the Lucky Imaging technique.
Left-hand panel is a simple average corresponding to a conventional long exposure. The middle and right-hand panels show images
resulting from re-centering and drizzling the short exposures with 100 percent and 5 percent selection cutoffs, respectively. All images
have the same log scale. The green bar depicts 1(cid:48)(cid:48)
line. Average seeing during these observations was 0.8(cid:48)(cid:48).
Table 1. The sample of 16 stars harbouring transiting extrasolar planets studied in this paper. We list the spectral type and the
orbital eccentricity (e) 2, λ the measured spin-orbit (mis)alignment angle (data and references taken from the Rossiter-McLaughlin
encyclopaedia1), the number of observed frames and the total exposure times for the 100% selection images of LuckyCam, and the
average seeing at 500nm.
Target
SpT
e
λ
HAT-P-1
HAT-P-2
HAT-P-5
HAT-P-6
HAT-P-8a
G0V 0.067
F8V
0.52
G1V
F8V
0
0
F
0
HAT-P-11
K4V
0.19
HD 209458 G0V
WASP-3
WASP-3
WASP-10
XO-1
F7V
"
K5V
G1V
0
0
"
0.05
0
(deg)
3.7 ± 2.1
−
1.2 ± 13.4/0.2+12.2−12.5/9 ± 10
166 ± 10/165 ± 6
−9.7+9.0−7.7/ − 17+9.2−11.5
103+22−18/103+26−10/106+15−11/97+8−4
3.9+18−21/ − 4.4 ± 1.4/ − 5 ± 7
13+9−7/3.3+2.5−4.4/5+6−5
"
−
−
Targets with candidate companions from this work
CoRoT-2b
CoRoT-3c
HAT-P-7d
TrES-1
TrES-2e
TrES-4e
G7V
F3V
F6V
K0V
G0V
F8V
0
0
0
0
0
0
7.2 ± 4.5/ − 1+6−7.7/4.7 ± 12.3
182.5 ± 9.4/ − 132.6+10.5−16.3/155 ± 37
−37.6+22.3−10
30 ± 21
−9 ± 12
6.3 ± 4.7
Nframes
Texp
(sec)
seeing
((cid:48)(cid:48))
5400
8000
9000
5000
9100
6000
10000
10000
5000
5000
10000
6000
5174
5175
7700
7000
10000
260
384
432
240
437
288
480
480
240
240
480
288
248
248
370
336
480
0.65
0.99
1.02
0.66
1.51
0.64
0.96
1.11
0.65
0.74
0.79
1.37
1.44
1.10
0.88
0.83
1.12
a Candidate companion identified by Bergfors et al. (2013)
b Candidate companion identified by Alonso et al. (2008)
c Candidate companion identified by Deleuil et al. (2008)
d Candidate companion identified by Narita et al. (2010)
e Candidate companion identified by Daemgen et al. (2009)
moval was applied. The LuckyCam pipeline registers the
image motion of each exposure using an interpolated cross-
correlation algorithm (Law et al. 2006; Staley & Mackay
2010). The peak of the cross-correlation map provides a
proxy for the Strehl ratio (i.e. the peak value of the PSF di-
vided by the theoretical diffraction-limited value, commonly
used as a high-resolution imaging performance metric) and
estimates the relative exposure quality (Staley & Mackay
2010). For each data set, re-centred and drizzled (Fruchter
& Hook 2002) images are produced by the pipeline which
then selects and co-adds observed frames that meet the
image quality criteria as described in detail by Law et al.
(2006); Staley & Mackay (2010). This procedure yields two
images for each data set, the first obtained by co-adding
the sharpest 5% selection of the frames, and the second by
co-adding all exposures (100%; see for example Figure 1 --
middle panel). When choosing the selection cutoff there is
a trade-off to be made between a smaller FWHM at low
percentage cutoffs (from fewer images with higher Strehl ra-
tio), and lower pixel noise at high percentage cutoffs, due
to longer cumulative exposure time. Figure 1 shows the im-
provement obtained with Lucky Imaging for the case of the
planet hosting star, TrES-2 (see also Law et al. 2006, Fig.2).
The NOT telescope is subject to aberrations and
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4
F. Faedi et al.
Figure 2. Different stages of the analysis of HAT-P-7. Top-left:
The image output from the LuckyCam pipeline with the 5% selec-
tion criterion. Bottom-left: After PSF subtraction, step 1 of the
image analysis algorithm (described in section § 4.2). Top-right:
After background subtraction via a median boxcar filter (Step 2).
Bottom-right: After Gaussian convolution (Step 3). The identified
close companion to HAT-P-7 is circled (North is left and East is
down.)
does not yield near diffraction-limited images (see e.g.
http://www.not.iac.es/telescope/tti/imqual.pdf). A combi-
nation of small-scale mirror irregularities and chromatic dis-
persion effects limit the probability of obtaining diffraction-
limited images although the large number of images and the
random phase variation of the atmosphere can compensate
for slight aberrations and telescope focusing.
Additionally, our Lucky Imaging data do not show
"quasi-static speckles", as in adaptive optics imaging (see
e.g. Marois et al. 2003; Boccaletti et al. 2003, 2004; Hink-
ley et al. 2007), that could be mistaken for faint compan-
ions. Our data were visually inspected in order to confirm
the presence of faint companion candidates throughout the
data reduction process. Furthermore, other possible causes
of false detection such as "ghosting" were not observed in
our data.
4.2 Image Analysis
The technique most widely applied when attempting to iden-
tify faint or crowded point sources in astronomical images
is that of PSF fitting and subtraction. A step crucial to this
process is the choice and evaluation of PSF models, which
may be derived semi-analytically (Dolphin 2000), empiri-
cally (Diolaiti et al. 2000), or by some combined analytical
model fit with empirical corrections (Stetson 1987). In the
case of Lucky Imaging we expect the PSF to be symmet-
ric, however it is not trivial to model the radial profile as
the PSF consists of a narrow core surrounded by a wide
halo (e.g., Hardy 1998). Our image analysis algorithm is de-
scribed below in three steps:
(1) PSF Subtraction: To create an axisymmetric,
semi-empirical model of the PSF we perform a Gaussian
fit of 9 pixels around the brightest, central pixel giving a
PSF central position to sub-pixel precision. The flux values
in the pixels around the nominal centre are collected into
bins (in radius) and a median and standard deviation are
evaluated at approximately one pixel-width radius intervals.
Any visually identified candidate in our images is masked
off during this process so as not to contaminate the PSF
model. The Gaussian fit is used within 1.5 pixels radius from
the PSF centre, while at larger radii the model is generated
using interpolated median values from the annulus bins.
Finally, the PSF model
is subtracted from the original
image to give a residual image shown in Fig. 2 (bottom-left).
(2) Background Subtraction via Median Boxcar
Filter: After the axisymmetric PSF model has been
subtracted, some artefacts can remain in the image that
might hamper attempts to identify companion stars. In
order to validate our detections, we employed a median
boxcar filter to suppress any artefacts present. For every
pixel, the background level
is estimated by taking the
median of all pixel values within a circular aperture of
radius 7 pixels (i.e. small enough to suppress localised
background variations, whilst remaining significantly larger
than the PSF core so that companion candidates are not
removed). The 'background map' of median values is then
subtracted from the residual image (see top-right, Fig. 2).
(3) Convolution with a Gaussian Profile: For this
relatively small dataset we visually inspected all the sources,
utilising a Gaussian convolution of the resulting images
from Step 2 to enhance visibility of any companion candi-
date (see bottom-right, Fig. 2). Once a candidate has been
re-identified, the location is inspected in images from all
stages of the image analysis process (i.e. reduced image, psf-
subtracted image, and background subtracted image) in or-
der to verify that the candidate is not a detector artefact or
arising from the image analysis process.
4.3 Candidate Detection
Our detection threshold was chosen to be 4 times the stan-
dard deviation of the background (σ) at any given concentric
circle at increasing separations from the centre of the planet
host. The sensitivity of our observations to detect stellar
companions at different angular separations from the pri-
mary planet host is given in Table 2. We place upper limits
in ∆i(cid:48) to the presence of stellar companions to all targets
at angular separations of 0.25(cid:48)(cid:48), 0.5(cid:48)(cid:48), 1(cid:48)(cid:48), 2.5(cid:48)(cid:48), and 6.5(cid:48)(cid:48) from
the centre of the primary. The adopted 4σ -- detection limits
depend on the exposure time and primary target magnitude
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Lucky Imaging of transiting planet hosts
5
Figure 3. The effect of seeing conditions on our detection sen-
sitivity for the planet host star WASP-3. The first 10,000 ex-
posures were obtained with an average seeing of 1.11(cid:48)(cid:48) and the
second 5,000 frames with an average seeing of 0.65(cid:48)(cid:48). The data
set obtained during better seeing conditions shows an increase in
detection sensitivity, important at small separations within the
seeing disc of the primary star.
as well as seeing. This is exemplified in Figure 3 where we
plot our detection sensitivity as a function of angular sepa-
ration in the case of WASP-3 during observations obtained
over two consecutive nights with different seeing conditions.
The first set of 10,000 images were obtained with an average
seeing of 1.11(cid:48)(cid:48) while the second 5,000 images were obtained
with an average seeing of 0.65(cid:48)(cid:48). The effect of poorer image
quality is particularly evident at small separations within
the seeing disc of the planet host star. Even though, the
first set of data have twice the number of frames, the images
taken during better seeing conditions allow the detection of
companions ∆i(cid:48) =1.8 magnitudes fainter at a separation of
0.25(cid:48)(cid:48). Figure 4 shows our average sensitivity. We depict our
results in black circles and our non-detections in red circles.
These are discussed in detail in Appendix A. Additionally
we report the minimum, average and maximum sensitivity
curves (grey dashed, dot-dashed lines) derived for the sam-
ple of host stars with no visually detected companions. Typi-
cally we can detect companions that are ∆i(cid:48) ∼ 4 magnitudes
fainter than the primary at a distance of 0.25(cid:48)(cid:48). As expected,
our sensitivity to fainter companions increases with increas-
ing distance from the planet host.
Once a candidate companion has been identified, it is
verified as a bona fide stellar source by excluding it as a
product of the data and/or image analysis as follows. First,
the FWHM of the planet host is measured from our im-
ages. Second, the flux of the primary star is measured us-
ing a circular aperture of diameter 6×FWHM on these im-
ages. Third, on the PSF subtracted images, a Gaussian fit
is used to determine the central pixel position of the candi-
date companion. Then, the flux of the identified companion
is measured on the PSF subtracted image, similarly to the
primary flux measurement. Finally, the signal-to-noise ratio
(SNR) of the companion (see Table 3) is calculated taking
into account background and photon shot noise using the
following equation:
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Figure 4. Average sensitivity of the LuckyCam survey (black
line). We indicate our detections with black circles and the three
non-detections discussed in Appendix A with red circles. The grey
dashed and the dot-dashed lines indicate our minimum, maximum
and average detection limits (see Table 2) for the sample of host
stars with no visual companion detected.
(cid:112)Npixσ2
F − Npixb
a + F
SN R =
(1)
where b and σ2
a are the mean value and variance of the back-
ground pixels within the aperture of the companion, F is the
flux over the number of pixels in the photometric aperture,
Npix. The SNR values for the candidate companions identi-
fied in this work are given in Table 3.
5 RESULTS
In the sample of 16 transiting planet host stars we have de-
tected candidate companion stars for six planet hosts TrES-
1, TrES-2, TrES-4, HAT-P-7, CoRoT-2 and CoRoT-3. Each
candidate companion has been identified from visual inspec-
tion of the reduced Lucky Imaging frames as described in
section § 4.3. We summarise our results in Table 3 where
we give the relative photometry and astrometry of the com-
panion candidates. To have a uniform dataset we only list
detections within 6.5(cid:48)(cid:48) from the centre of the planet host
star.
Our LuckyCam images clearly show the presence of two
candidate companions to the planet hosts star TrES-1, pre-
viously unknown. Figures 5 shows the LuckyCam images
for TrES-1 and the candidate companions identified in this
work.
Among our sample TrES-2, TrES-4 and HAT-P-7 have
previously published high-resolution Adaptive Optics (AO)
and/or Lucky Imaging observations showing the presence
of faint stellar companions (Daemgen et al. 2009; Narita
et al. 2010; Bergfors et al. 2013). Additionally, the com-
panion stars to CoRoT-3 (2MASS J19281330+0007135) and
CoRoT-2 (2MASS J19270636+0122577), have been iden-
tified in previous works see e.g. Deleuil et al. (2008) and
Alonso et al. (2008), Gillon et al. (2010), respectively. We
note that Deleuil et al. (2008) also mentions a second fainter
companion to CoRoT-3 at separation 5.6(cid:48)(cid:48) . We do not re-
port this object in our discussion as it falls near the CCD
edge making our image analysis unreliable. The compan-
6
F. Faedi et al.
Table 2. The 4σ-detection limits (in ∆i(cid:48)) for all stars in our
sample with and without detected companions at separations of
r = 0.25(cid:48)(cid:48), 0.5(cid:48)(cid:48), 1(cid:48)(cid:48), 2.5(cid:48)(cid:48), and 6.5(cid:48)(cid:48).
4σ -- detection limits (∆i(cid:48))
0.25
0.5
1.0
2.5
6.5
r((cid:48)(cid:48))
Target
HAT-P-1
HAT-P-2
HAT-P-5
HAT-P-6
HAT-P-8
HAT-P-11
HD 209458
WASP-10
WASP-3b
WASP-3a
XO-1
3.57
4.17
3.81
3.63
4.67
3.29
4.32
3.86
4.36
2.66
3.39
4.12
4.51
4.05
4.78
4.73
4.54
4.42
4.38
4.03
4.31
4.03
7.62
6.21
6.04
7.50
5.75
7.58
7.12
6.39
7.08
5.73
6.80
5.36
5.51
5.55
7.03
6.82
6.15
8.99
8.16
7.50
8.87
7.52
9.51
9.24
7.31
8.78
7.71
8.05
6.26
5.92
7.43
7.97
7.50
7.86
9.22
8.23
7.70
9.12
7.97
9.72
9.31
7.41
9.18
8.16
8.19
6.37
5.89
7.91
8.02
7.50
7.99
Targets with candidate companions
CoRoT-2
CoRoT-3
HAT-P-7
TrES-1
TrES-2
TrES-4
4.43
4.90
4.14
4.11
4.40
4.19
4.71
5.09
4.51
4.52
5.14
4.26
a Derived from 10,000, compared to b 5,000.
Figure 5. The LuckyCam images for the planet host star TrES-
1. North is left and East is down. The two companions are clearly
visible in our images.
ion to CoRoT-2 and the two companions to HAT-P-7 have
also been confirmed to be bound to the planet-hosting stars,
forming wide binary systems (Schroter et al. 2011; Narita
et al. 2012). We confirm previous findings for the com-
panions to TrES-2 and TrES-4, while for HAT-P-7 we can
only detect the brighter of the two companions found by
Narita et al. (2010, 2012). The authors estimated the fainter
companion to HAT-P-7 to be of spectral type M9 − L0
(m2 (cid:39) 0.078 − 0.088 M(cid:12)) at a separation of 3.14 ± 0.01(cid:48)(cid:48),
and therefore is below our detection limit for these observa-
tions, see Tables 2 and 4. Our results for the position angles,
spectral type determinations, and separations for the com-
panions to CoRoT-2, TrES-2, TrES-4 and HAT-P-7 agree
Table 3. Results for the planet hosting stars with detected candi-
date stellar companions from this work. From left to right we list
the name of the planet host, the angular separation of the candi-
date companion, the position angle, the ∆i(cid:48) magnitude, and the
SNR of the detected companion.
Target
CoRoT-2
CoRoT-3
HAT-P-7
TrES-1
TrES-1
TrES-2
TrES-4
r
((cid:48)(cid:48))
4.10 ± 0.03
5.24 ± 0.03
3.87 ± 0.03
4.95 ± 0.03
6.19 ± 0.03
1.11 ± 0.03
1.54 ± 0.03
PA
(◦)
208.4 ± 0.4
173.9 ± 0.4
90.4 ± 0.5
149.6 ± 0.5
47.4 ± 0.2
137 ± 2
1.2 ± 1.2
∆i(cid:48)
(mag)
2.95 ± 0.03
3.0 ± 0.1
6.9 ± 0.1
6.02 ± 0.08
5.79 ± 0.07
3.97 ± 0.01
4.51 ± 0.02
SNR
41
10
10
14
17
86
52
with the results obtained by Alonso et al. (2008); Gillon
et al. (2010); Schroter et al. (2011), Daemgen et al. (2009),
Narita et al. (2010, 2012), and Bergfors et al. (2013), and are
presented in Tables 5 and 6. In the case of the planet host
star HAT-P-8 we are unable to confirm the candidate com-
panion identified by Bergfors et al. (2013). The sensitivity
of our observations of HAT-P-8 at the separation of 1.027(cid:48)(cid:48)
would only allow us to detect companions two magnitudes
brighter than the detection reported by the authors.
5.1 Non-Detections
Our visual inspection of the LuckyCam images showed no
stellar companions to the following planet host stars: HAT-
P-1, HAT-P-2, HAT-P-5, HAT-P-6, HAT-P-8, HAT-P-11,
HD 209458, WASP-10, WASP-3, and XO-1. Our results are
in agreement with previous studies with the exception of
HAT-P-8 for which our reduced image quality does not al-
low the identification of the companion reported by Bergfors
et al. (2013). Finally, in the cases of HD 209458, HAT-P-
5, and HAT-P-6 our visual inspection of the images shows
possible candidate companions to the planet hosts, however,
after further consideration (discussed in appendix A), these
putative identifications are classified as non-detections.
6 STATISTICAL LIKELIHOOD OF
ASSOCIATION
The detection of faint stellar companions associated with our
targets could provide important observational constraints
for theoretical models of planet formation and evolution. We
used a statistical approach to investigate the probability of
each detected companion star being gravitationally bound to
the planet host. We first estimated the density of background
sources ρ(m) in a cone of 10(cid:48) around each target. Because
our targets are quite bright we used the 2MASS catalogue to
retrieve objects within the 10(cid:48) cone around the planet host
coordinates. Subsequently, we derive the probability that a
target star has a non-related background source within the
separation of the detected candidate companions. By using
a similar method to that adopted by Daemgen et al. (2009),
we used the 2MASS magnitudes to identify bright giant stars
in the ensemble of retrieved objects. We selected all objects
with J − Ks > 0.5 and with K < 15, which corresponds to
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
the background detection limit in these short accumulated
exposures (see Texp in Table 1)3. There is a degeneracy in
the near-IR colours of giant and dwarf stars for early spectral
types (earlier than K7 or J − Ks > 0.5), but these become
distinct in two-colour diagrams for the latest spectral types
(Majewski et al. 2003). Juri´c et al. (2008) used a model
of our Galaxy to estimate the number of giant stars which
could be misidentified as main-sequence stars and found that
the overall bias in the estimated number density is ∼ 4%
within 500 pc. Finally, we used Equation 1 from Brandner
et al. (2000) to find the probability P(Θ, m) for an unrelated
source to be located within a certain angular distance Θ from
the target.
P(Θ, m) = 1 − e
−πρ(m)Θ2
(2)
where Θ is in arcsec and ρ(m) is the estimated density of
background sources within 10(cid:48) of the target. We calculated
P(Θ, m) for each star with a detected faint companion can-
didate. We also used our images to estimate the expected
number of sources in our images with background, not as-
sociated, companions (see column 6 of Table 5). We note
that all but the CoRoT targets have a very low probability
of contamination by background sources (see Table 5). The
CoRoT satellite observes alternatively towards the galac-
tic centre and anti-centre, thus increasing the probability of
contamination by background objects.
To further test the probability of chance alignment for
the binary pairs we used an independent statistical analy-
sis following the method described in (Dhital et al. 2010).
We calculated the frequency of unrelated pairings using a
Galactic model that is parameterised by an empirically mea-
sured stellar number density distribution in a 30(cid:48) × 30(cid:48) con-
ical volume centred on the candidate binary. The simulated
stellar distributions are constrained by empirical measure-
ments from the Sloan Digital Sky Survey (Juri´c et al. 2008;
Bochanski et al. 2010) and accurately accounts for the de-
crease in stellar number density with both galactocentric ra-
dius and galactic height. All the simulated stars are, by def-
inition, single and unrelated. Therefore, the total the num-
ber of simulated stars that are nearby to the candidate pri-
mary is the likelihood that the candidate binary is a chance
alignment. We performed 106 realisations for each of our six
candidate binaries. Table 5, column 6 and 8, show both es-
timated probabilities P(Θ, m) and PD10, respectively. Our
results strongly suggest that all the detected faint compan-
ions within 6.5(cid:48)(cid:48) to our targets are not random chance align-
ments.
6.1 Companion Properties
Under the assumption that the detected companions are
bound to the planet host stars in our sample, we used
2MASS magnitudes, spectral types, and temperatures (Teff )
of the planet-host targets to derive spectral types and masses
for each candidate companion discussed in Section §5. We
first estimated absolute MJ, MH, MK magnitudes for each
planet-host star using their published distances and 2MASS
3 We note however, that for bright guide stars longer observations
would have allowed the detection of background sources as faint
as i(cid:48) ∼ 22 (see e.g. Law et al. 2006)
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Lucky Imaging of transiting planet hosts
7
magnitudes. We then used the models given in Table 5 of
Kraus & Hillenbrand (2007), and models from Baraffe et al.
(cid:48)
(1998) to evaluate the absolute i
magnitude for the planet
hosts interpolating within MJ, MH, MK, and Teff . In Ta-
ble 5 we give the estimated Mi(cid:48) , spectral types and masses
for each candidate companion. We note that the candidate
companions identified in this study have spectral types later
than K4 (see also Daemgen et al. 2009; Bergfors et al. 2013;
Narita et al. 2010; Schroter et al. 2011), making them dif-
ficult to identify in optical spectra, as well as in optical,
seeing-limited photometry.
The faint stellar companions identified to TrES-1 have
separations from it larger than 2(cid:48)(cid:48), sufficient to avoid blend-
ing effects during spectroscopic and photometric observa-
tions. Such effects in the case of TrES-2, TrES-4 and HAT-
P-7 have been investigated by Daemgen et al. (2009) and
Bergfors et al. (2013) and have been found to be not sig-
nificant. Under the assumptions above, we derived physical
separations, spectral types and masses for the companions
to TrES-2, TrES-4 and HAT-P-7 that are in agreement with
previous results (see Tables 5 and 6). For the companion to
CoRoT-2 the 2MASS magnitudes are J = 12.866 ± 0.033, H
= 12.234 ± 0.044, and K = 12.028 ± 0.031. Using the pub-
lished distance of CoRoT-2 and the models from Kraus &
Hillenbrand (2007) we obtain a spectral type of M0 (±1
SpT), in agreement with the estimate by Schroter et al.
(2011).
The candidate companion to CoRoT-3 is also visi-
ble in 2MASS images Cutri et al. 2003, and both stars
are classified as 2MASS J19281330+0007135 and 2MASS
19281326+0007185, respectively. The near-IR magnitudes
of CoRoT-3 are J = 14.027 ± 0.036, H = 13.448 ± 0.045,
and K = 13.295 ± 0.043. The separation between the ob-
jects given in the 2MASS catalogue is 5.1± 0.1(cid:48)(cid:48), in position
angle 173◦, which are in good agreement with the value of
5.24 ± 0.03(cid:48)(cid:48) obtained in this work. Our chance alignment
probability for CoRoT-3 is the highest amongst the values
derived in this work, however, the proper motions from the
NOMAD catalogue (Zacharias et al. 2005) for CoRoT-3 are
µα = −10.7 ± 5.6mas/yr and µδ = 21.8mas/yr, which over
the 9 yr span between the 2MASS and our observations give
a total proper motion of about 0.2(cid:48)(cid:48). Therefore, our results
are consistent with the candidate companion being bound
to CoRoT-3. Assuming the object is at the same distance as
CoRoT-3 we derive a spectral type of K4 -- K5 (see Table 4).
7 SUMMARY
To date several different hypotheses have been formulated
in order to explain the observed properties of planetary sys-
tems. Compared to our own solar system, gas giant planets
have been found with very short period orbits (P< 10d)
posing the problem and at the same time, providing evi-
dence of planetary migration (Lin et al. 1996, Wu & Mur-
ray 2003, Ida & Lin 2004, Nagasawa et al. 2008, Marzari
& Nelson 2009). The existence of giant planets in highly
eccentric orbits and the measurements of their spin-orbit
(mis)alignments demonstrate that there must be a number
of mechanisms capable of shaping the system orbital config-
uration. Although evidence for such mechanisms has been
provided (Winn et al. 2010, Triaud et al. 2010, H´ebrard et al.
8
F. Faedi et al.
2010, Narita et al. 2010, Schlaufman 2010), it is not yet clear
which specific mechanisms are more important or act at a
particular time to sculpt the configuration of known plane-
tary systems. Recently Albrecht et al. (2012) suggested that
the Kozai mechanism is responsible for the migration of the
majority, if not all, hot Jupiters, those mis-aligned as well
as those aligned, and that star-planet tidal interaction plays
a central role in shaping exoplanets orbital configurations.
Moreover, Narita et al. (2012) suggest that the presence of
the two bound companion stars to HAT-P-7 can provide an
explanation of the planetary mis-aligned orbit via sequential
Kozai migration (Takeda et al. 2008). Thus, the detection of
faint companions to the planet hosts will provide important
observational evidence, fundamental for the understanding
of the formation and evolution of their planetary systems.
We have investigated the presence of faint stellar com-
panions within 6.5(cid:48)(cid:48) of 16 host stars of transiting exoplanets
by means of the Lucky Imaging technique. We show that this
technique has the potential to detect faint stellar compan-
ions within the seeing disc (< 1(cid:48)(cid:48)) of bright primary stars.
We have identified faint candidate stellar companions
to six planet hosts. Over the range of brightness of the se-
(cid:48) < 5.47,
lected planet host stars in our sample (3.50 < Mi
i.e. 7.65 < V < 14) we give 4-σ detection limits for puta-
tive companions at increasing separations of 0.25(cid:48)(cid:48), 0.5(cid:48)(cid:48), 1(cid:48)(cid:48),
2.5(cid:48)(cid:48) and 6.5(cid:48)(cid:48) from the centre of the primary. For the targets
with no detections, we are able to exclude stellar compan-
ions of spectral types between M1 and M8 at separations
> 1(cid:48)(cid:48), depending on the brightness of the primary and the
seeing at which the object was observed (see Fig. 4).
We have identified two faint candidate companions to
the planet host TrES-1 that have not been previously re-
ported, and our statistical analysis suggests that these stars
could be bound to the planet host. Assuming that all the
candidate companions are bound to the planet hosting stars,
we used the known distances together with models from
Kraus & Hillenbrand (2007), and models from Baraffe et al.
(1998) to estimate spectral types and masses. In the case of
TrES-1 we find the first companion at separation 4.95±0.03(cid:48)(cid:48)
to be of spectral type M5 (±1SpT) implying a mass of
0.15M(cid:12). The second at separation of 6.19±0.03(cid:48)(cid:48) is found to
be of spectral type M5 and mass between 0.2 and 0.15M(cid:12).
In the case of CoRoT-3 we obtain a spectral type of K4 -- K5
and a stellar mass between 0.75 and 0.7M(cid:12) for the candidate
companion. For TrES-2, TrES-4, HAT-P-7, and CoRoT-2 we
confirm both known candidates as well as bound compan-
ions and our estimated spectral types and masses agree with
those found by Daemgen et al. (2009); Bergfors et al. (2013);
Narita et al. (2010) and Schroter et al. (2011); Narita et al.
(2012). Overall, for our targets the epoch of observations ei-
ther coincide with that of previous works (e.g. Bergfors et al.
2013; Narita et al. 2012), or only allow a short temporal
separation with respect to archival and published observa-
tions. Given the precision of our astrometry and the relative
proper motions of the target stars this does not allow any ro-
bust conclusion on the binarity of the detected companions.
Therefore, additional high-resolution high-contrast imaging
observations are necessary in order to robustly confirm if the
companions observed in this and previous works are bound
the the planet host stars.
Finally, we discuss in an Appendix the cases of HD
209458, HAT-P-5 and HAT-P-6, for which possible stellar
companions were initially visual identified in our images but
subsequently classified as non-detections after further anal-
ysis was carried out.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Table 4. Upper limits for companions' spectral types and masses for the 16 planet host stars in our sample at separations r =
0.25(cid:48)(cid:48), 0.5(cid:48)(cid:48), 1(cid:48)(cid:48), 2.5(cid:48)(cid:48), and 6.5(cid:48)(cid:48) from the primary.
r = 0.25(cid:48)(cid:48)
r = 0.5(cid:48)(cid:48)
r = 1(cid:48)(cid:48)
r = 2.5(cid:48)(cid:48)
r = 6.5(cid:48)(cid:48)
Lucky Imaging of transiting planet hosts
9
Target
Sp.T
±1
m2
M(cid:12)
Sp.T
±1
m2
M(cid:12)
Sp.T
±1
m2
M(cid:12)
0.15
0.20 -- 0.15
0.20 -- 0.15
0.15
0.20
M5
0.54 -- 0.42
0.54 -- 0.42 M4 -- M5
M4 -- M5
M5
M4
0.54
0.42
0.42
0.15
0.42
0.54
M4 -- M5
0.20 -- 0.15
M4
0.59 -- 0.54
0.20 -- 0.15 M6 -- M7
0.54 -- 0.42
M5
0.20
0.12 -- 0.11
Sp.T
±1
M7
M7
M6
M6
M6
L0
m2
M(cid:12)
0.11
0.11
0.12
0.12
0.12
0.078
M6
M7
M8
0.12
0.11
0.102
Sp.T
±1
M7
M7
M6
M7
M6
>L0
M7
M6
M7
M8
m2
M(cid:12)
0.11
0.11
0.12
0.11
0.12
<0.078
0.11
0.12
0.11
0.102
0.15
M6 -- M7
0.12 -- 0.11 M6 -- M7
0.12 -- 0.11
M0
M0
M2
M2
K7 -- M0
M0 -- M1
HAT-P-1
HAT-P-2
HAT-P-5
HAT-P-6
HAT-P-8
HAT-P-11
HD 209458 M1 -- M2
WASP-3a
WASP-3b
WASP-10
XO-1
K5
M0
M4
K7
0.63 -- 0.59 M1 -- M2
0.59 -- 0.54 M1 -- M2
0.59
0.59
0.42
0.42
0.54 -- 0.42
0.70
0.59
0.20
0.63
M1
M2
M2
M5
M2
M1
M0 -- M1
M4 -- M5
M1 -- M2
Targets with companion candidates
CoRoT-2
CoRoT-3
HAT-P-7
TrES-1
TrES-2
TrES-4
M2 -- M3
0.42 -- 0.29
M2
M1
M2 -- M3
M1 -- M2
M0
0.42
0.54
0.59
0.42 -- 0.29 M3 -- M4
0.54 -- 0.42
M3
M1
M1
M3
M1
M6 -- M7
0.12 -- 0.11
M5
0.15
M6 -- M7
0.12 -- 0.11
0.29
0.54
0.54
0.29 -- 0.20
0.29
0.54
M4
M2
M3 -- M4
M6
M5
0.29 -- 0.20 M5 -- M6
M7 -- M8
M6 -- M7
0.12
0.15
M4 -- M5
0.20 -- 0.15
M6
0.20
0.42
M4 -- M5
0.20 -- 0.15
M3
0.29
M5
M3
M6
M7
0.15 -- 0.12
0.11 -- 0.102
0.12 -- 0.11 M6 -- M7
M6 -- M7
0.12
0.15
0.29
0.12
0.11
0.12 -- 0.11
0.12 -- 0.11
Table 5. Companion candidates for 6 planet host stars. From left to right we list the name of the planet host star, separation angle,
the position angle, the ∆i(cid:48) for the detected companions, the SNR of the detected companion, the probability for the companion to be a
chance alignment (P(Θ, m)) and the expected number of sources with an unrelated background companion (Ebg), the probability of a
chance alignment detection as estimated by Dhital et al. (2010), the planet host's distance (pc), and finally the companion separation in
AU, assuming the value is a lower limit.
Target
CoRoT-2
CoRoT-3
HAT-P-7
TrES-1
TrES-1
TrES-2
TrES-4
r
((cid:48)(cid:48))
4.10 ± 0.03
5.24 ± 0.03
3.87 ± 0.03
4.95 ± 0.03
6.19 ± 0.03
1.11 ± 0.03
1.54 ± 0.03
PA
(◦)
208.4 ± 0.4
173.9 ± 0.4
90.4 ± 0.5
149.6 ± 0.5
47.4 ± 0.2
137 ± 2
1.2 ± 1.2
∆i(cid:48)
(mag)
2.95 ± 0.03
3.0 ± 0.1
6.9 ± 0.1
6.02 ± 0.08
5.79 ± 0.07
3.97 ± 0.01
4.51 ± 0.02
SNR P(Θ, m)
Ebg
(%)
3.17
4.05
0.03
0.82
1.29
0.03
0.03
41
10
10
14
17
86
52
0.22
0.12
0.004
0.025
0.039
0.0005
0.0007
PD10
(%)
1.18
1.72
0.2
0.04
0.06
0
0
Dist.
(pc)
270 ± 120
680 ± 160
320 ± 50
150 ± 6
150 ± 6
220 ± 10
479 ± 26
Sep.
(AU)
1108 ± 492
3562 ± 838
1238 ± 193
743 ± 30
929 ± 37
244 ± 13
740 ± 43
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
10
F. Faedi et al.
e
r
a
e
r
a
s
n
o
i
n
a
p
m
o
c
e
h
t
r
o
f
s
e
u
l
a
V
.
n
o
i
n
a
p
m
o
c
h
c
a
e
r
o
f
y
t
i
r
a
n
b
i
i
g
n
m
u
s
s
a
d
e
v
i
r
e
d
,
s
r
a
t
s
n
o
i
n
a
p
m
o
c
e
h
t
r
o
f
s
e
s
s
a
m
d
n
a
s
e
p
y
t
l
a
r
t
c
e
p
s
,
)
(cid:48)i
M
(
s
e
d
u
t
i
n
g
a
m
(cid:48)
i
e
t
u
l
o
s
b
a
d
e
t
a
m
i
t
s
E
.
6
e
l
b
a
T
s
r
o
r
r
e
e
d
u
t
i
n
g
a
M
.
s
t
e
g
r
a
t
t
s
o
h
t
e
n
a
l
p
e
h
t
f
o
ff
e
T
d
n
a
s
e
c
n
a
t
s
i
d
,
s
e
d
u
t
i
n
g
a
m
S
S
A
M
2
d
e
h
s
i
l
b
u
p
g
n
i
s
u
s
l
e
d
o
m
)
8
9
9
1
(
.
l
a
t
e
e
ff
a
r
a
B
d
n
a
)
7
0
0
2
(
d
n
a
r
b
n
e
l
l
i
H
&
s
u
a
r
K
m
o
r
f
d
e
v
i
r
e
d
.
y
l
e
v
i
t
c
e
p
s
e
r
,
)
s
(
n
o
i
n
a
p
m
o
c
e
h
t
d
n
a
r
a
t
s
t
s
o
h
e
h
t
e
t
a
c
i
d
n
i
2
d
n
a
1
t
p
i
r
c
s
r
e
p
u
S
.
s
e
c
n
a
t
s
i
d
d
n
a
e
d
u
t
i
n
g
a
m
K
,
H
,
J
t
e
g
r
a
t
e
h
t
n
o
s
r
o
r
r
e
n
w
o
n
k
e
h
t
f
o
n
o
i
t
a
g
a
p
o
r
p
h
g
u
o
r
h
t
d
e
t
a
m
i
t
s
e
2
m
)
(cid:12)
M
(
9
5
.
0
0
7
.
0
--
5
7
.
0
5
1
.
0
--
0
2
.
0
5
1
.
0
5
1
.
0
--
0
2
.
0
2
4
.
0
--
4
5
.
0
2
4
.
0
2
T
p
S
(cid:48)
2 i
M
0
M
5
K
--
4
K
5
M
--
4
M
5
M
5
M
2
M
--
1
M
2
M
5
5
.
7
5
4
.
6
2
9
.
0
1
9
4
.
1
1
7
2
.
1
1
1
4
.
8
6
7
.
8
(cid:48)
i
∆
)
g
a
m
(
5
9
.
2
0
0
.
3
2
9
.
6
2
0
.
6
0
8
.
5
7
9
.
3
0
5
.
4
2
1
.
0
--
5
1
.
0
6
M
--
5
M
2
1
.
0
0
1
.
0
6
M
8
M
--
7
M
7
9
.
1
1
1
3
.
2
1
8
3
.
4
1
7
5
.
7
1
9
.
7
9
6
.
0
1
(cid:48)
1 i
M
0
6
.
4
0
5
.
3
0
0
.
4
7
4
.
5
7
4
.
5
4
4
.
4
6
2
.
4
0
4
.
4
0
4
.
4
9
6
.
3
f
f
1e
T
)
K
(
1
T
p
S
1K
M
1H
M
1J
M
1
K
1
H
1
J
t
e
g
r
a
T
7
3
±
8
0
6
5
0
4
1
±
0
4
7
6
0
8
±
0
5
3
6
3
2
±
4
1
2
5
3
2
±
4
1
2
5
0
5
±
0
5
8
5
5
7
±
0
0
2
6
3
3
±
5
7
0
6
0
0
1
±
0
6
9
5
0
8
±
0
7
5
6
7
G
3
F
6
F
0
K
0
K
0
G
8
F
0
G
1
G
8
F
2
2
.
0
±
5
1
.
3
2
2
.
0
±
6
4
.
2
9
2
.
0
±
1
8
.
1
5
2
.
2
±
4
9
.
3
1
2
.
1
±
4
9
.
3
0
8
.
0
±
3
1
.
3
0
4
.
0
±
3
9
.
1
2
2
.
0
±
8
2
.
3
2
2
.
0
±
5
5
.
2
9
2
.
0
±
2
8
.
1
6
2
.
2
±
1
0
.
4
2
2
.
1
±
1
0
.
4
1
8
.
0
±
1
2
.
3
0
4
.
0
±
5
9
.
1
6
4
.
0
±
4
9
.
2
5
2
.
0
±
1
8
.
2
6
2
.
0
±
3
2
.
2
6
4
.
0
±
0
0
.
3
6
2
.
0
±
5
8
.
2
7
2
.
0
±
6
3
.
2
0
2
.
0
±
3
6
.
3
2
2
.
0
±
7
7
.
2
9
2
.
0
±
3
0
.
2
5
2
.
2
±
1
4
.
4
1
2
.
1
±
1
4
.
4
1
8
.
0
±
2
5
.
3
0
4
.
0
±
8
1
.
2
6
4
.
0
±
3
2
.
3
5
2
.
0
±
7
1
.
3
5
2
.
0
±
7
4
.
2
9
1
0
.
0
±
1
3
.
0
1
9
1
0
.
0
±
2
6
.
1
1
0
2
0
.
0
±
3
3
.
9
0
3
0
.
0
±
2
8
.
9
0
3
0
.
0
±
2
8
.
9
0
2
0
.
0
±
5
8
.
9
0
2
0
.
0
±
3
3
.
0
1
2
0
.
0
±
4
4
.
0
1
2
0
.
0
±
1
7
.
1
1
2
0
.
0
±
4
3
.
9
4
0
.
0
±
9
8
.
9
4
0
.
0
±
9
8
.
9
3
0
.
0
±
3
9
.
9
2
0
.
0
±
5
3
.
0
1
2
0
.
0
±
8
7
.
0
1
2
0
.
0
±
4
9
.
1
1
2
0
.
0
±
5
5
.
9
3
0
.
0
±
9
2
.
0
1
3
0
.
0
±
9
2
.
0
1
3
0
.
0
±
3
2
.
0
1
2
0
.
0
±
8
5
.
0
1
†
2
-
T
o
R
o
C
†
3
-
T
o
R
o
C
-
7
-
P
T
A
H
1
-
S
E
r
T
1
-
S
E
r
T
2
-
S
E
r
T
4
-
S
E
r
T
0
2
0
.
0
±
1
3
.
6
0
2
0
.
0
±
8
4
.
0
1
0
3
0
.
0
±
1
3
.
9
2
0
.
0
±
7
3
.
6
3
0
.
0
±
2
5
.
0
1
4
0
.
0
±
4
4
.
9
2
0
.
0
±
9
5
.
6
2
0
.
0
±
4
8
.
0
1
2
0
.
0
±
6
5
.
9
8
5
4
9
0
2
D
H
-
5
-
P
T
A
H
-
6
-
P
T
A
H
s
n
o
i
t
c
e
t
e
d
-
n
o
N
s
n
o
i
n
a
p
m
o
c
d
n
u
o
b
d
e
m
r
fi
n
o
C
†
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
ACKNOWLEDGEMENTS
Based on observations made with the Nordic Optical Tele-
scope, operated on the island of La Palma jointly by Den-
mark, Finland, Iceland, Norway, and Sweden, in the Span-
ish Observatorio del Roque de los Muchachos of the Insti-
tuto de Astrofisica de Canarias. FF would like to thank
STFC for support through the award of a PDRA as part
of the QUB Rolling Grant for Astrophysics. Y.G.M.C. ac-
knowledges postdoctoral funding support from the Vander-
bilt Office of the Provost, through the Vanderbilt Initiative
in Data-intensive Astrophysics (VIDA) and through a grant
from the Vanderbilt International Office in support of the
Vanderbilt-Warwick Exoplanets Collaboration.
APPENDIX A: NON DETECTIONS WITH
VISUAL IDENTIFICATIONS
(i) HD 209458: For the planet host star HD 209458,
slight aberration effects are evident in our images resulting
from small scale mirror irregularities of the NOT, and
chromatic dispersion effects (Law et al. 2006). These effects
are more pronounced in the images of bright targets like HD
209458 (V=7.63; Høg et al. 2000). The possible detection
was present in all four stages of our image analysis at a
separation of 1.66(cid:48)(cid:48) (within the seeing disc of the primary
star) and position angle 241 ± 1◦ with ∆i(cid:48) = 7.57 (SNR
∼20). Figure A1 presents the PSF subtraction and the
Gaussian convolution steps for HD 209458 showing evidence
of the non-axisymmetric PSF, and of the possible detection.
Figure A2 shows our sensitivity as a function of separation
from the centre of the primary target. Our possible detec-
tion is well above our sensitivity limit at the separation
of 1.66(cid:48)(cid:48). However, any identification in our images within
the seeing disc of the planet host is investigated further for
possible artefacts. VLT+NACO images in the H-band for
HD 209458 are publicly available from the ESO archive4.
Our analysis of these NACO near-infrared adaptive optic
data do not show any evidence of a stellar companion at the
position of our possible detection. We would have expected
any stellar companion to be brighter in the near-infrared,
and thus be readily identifiable in the NACO photometry.
This is also in agreement with the non detection in the
Lucky Imaging observations by Daemgen et al. (2009)
and Bergfors et al. (2013). Therefore, we conclude that
the possible detection is most likely spurious due to the
limited image quality for HD 209458, resulting from the
seeing conditions, the number of frames, and the optical
characteristics of the NOT.
(ii) HAT-P-5: During our image analysis procedure
and visual
inspection of the images for the planet host
HAT-P-5 we have identified a candidate companion with
∆i(cid:48) = 7.9 (SNR∼1.9) at a separation of 4.25(cid:48)(cid:48) from the
centre of the primary star and position angle 268.5 ± 0.4◦.
Figure A3 shows the image from the 5% best LuckyCam
frames for HAT-P-5 (left), and the Step 3 (right) of the
image analysis where the candidate companion is clearly
visible. The measured ∆i(cid:48) is 0.14 magnitude below our 4-σ
4 ESO Archive: http://archive.eso.org/cms.html
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Lucky Imaging of transiting planet hosts
11
Figure A1. Spurious detection for HD 209458 at a separation of
1.66(cid:48)(cid:48) with a ∆i(cid:48) = 7.57. Left panel: Image after PSF subtraction
(Step 1 of image analysis). Right panel: Image after Gaussian
convolution (Step 3). The green cross marks the centre of the
primary star, whereas circled in green is the spurious detection
most likely due to our image quality.
Figure A2. Sensitivity curve as function of distance from the
primary plant host star HD 209458 derived for the 5% best-frame
selection. The vertical and horizontal grey-solid lines indicate the
angular separation (r = 1.66(cid:48)(cid:48)), and ∆i(cid:48) = 5.57 of the possible
detection, respectively. Our sensitivity at the angular separation
of 1.66(cid:48)(cid:48) is ∆i(cid:48) = 8.9.
Figure A3. Non-detection for HAT-P-5. Left panel: LuckyCam
5%-frame selection image for HAT-P-5 (Step 1). Right panel: The
Gaussian convolution image (Step 3). The green cross marks the
location of the centre of the primary star, the tentative companion
is circled in green.
12
F. Faedi et al.
Figure A4. Non-detection for HAT-P-6. Left panel: LuckyCam
5%-frame selection image for HAT-P-6 (Step 1). Right panel: The
Gaussian convolution image (Step 3). The green cross marks the
location of the centre of the primary star, the tentative companion
is circled in green.
detection cut-off at the separation of 4.25(cid:48)(cid:48), thus it was
classified as a non-detection.
(iii) HAT-P-6: In the images of the planet host HAT-P-
6, a candidate companion with ∆i(cid:48) ≈ 10.7 (corresponding
to a SNR ∼0.4) at a separation of 6.4(cid:48)(cid:48) is identified by visual
inspection. For example, Figure A4 shows the LuckyCam
images for Step 1 (left) and Step 3 (right) of the image
analysis where the candidate companion is clearly visible.
However, the measured ∆i(cid:48) of the putative companion is
more than one magnitude below the 4-σ detection threshold
at that separation from the centre of the primary (see
Table 2), thus it is considered a non-detection.
In the case of HAT-P-5 and HAT-P-6 our image sensitivity
does not allow us to reliably detect the putative companions.
However, because our images clearly show the presence of
possible companions at large separations from the primary,
these might be real and worth further investigation.
REFERENCES
Albrecht S., Winn J. N., Johnson J. A., Howard A. W.,
Marcy G. W., Butler R. P., Arriagada P., Crane J. D.,
Shectman S. A., Thompson I. B., Hirano T., Bakos G.,
Hartman J. D., 2012, ApJ, 757, 18
Alonso R., Auvergne M., Baglin A., Ollivier M., Moutou
C., Rouan D., Deeg H. J., et al., 2008, A&A, 482, L21
Anderson D. R., Hellier C., Gillon M., et al., 2010, ApJ,
709, 159
Bakos G. ´A., Howard A. W., Noyes R. W., et al., 2009,
ApJ, 707, 446
Bakos G. ´A., Noyes R. W., Kov´acs G., et al., 2007, ApJ,
656, 552
Bochanski J. J., Hawley S. L., Covey K. R., West A. A.,
Reid I. N., Golimowski D. A., Ivezi´c Z., 2010, AJ, 139,
2679
Brandner W., Zinnecker H., Alcal´a J. M., Allard F., Covino
E., Frink S., Kohler R., Kunkel M., Moneti A., Schweitzer
A., 2000, AJ, 120, 950
Burrows A., Hubeny I., Budaj J., Hubbard W. B., 2007,
ApJ, 661, 502
Chatterjee S., Ford E. B., Matsumura S., Rasio F. A., 2008,
ApJ, 686, 580
Crida A., S´andor Z., Kley W., 2008, A&A, 483, 325
Crossfield I. J. M., Barman T., Hansen B. M. S., Tanaka
I., Kodama T., 2012, ApJ, 760, 140
Cutri R. M., Skrutskie M. F., van Dyk S., et al., 2003,
VizieR Online Data Catalog, 2246, 0
Daemgen S., Hormuth F., Brandner W., Bergfors C., Jan-
son M., Hippler S., Henning T., 2009, A&A, 498, 567
Deleuil M., Deeg H. J., Alonso R., Bouchy F., Rouan D.,
Auvergne M., Baglin A., et al., 2008, aap, 491, 889
Dhital S., West A. A., Stassun K. G., Bochanski J. J., 2010,
AJ, 139, 2566
Diolaiti E., Bendinelli O., Bonaccini D., Close L., Currie
D., Parmeggiani G., 2000, A&AS, 147, 335
Dolphin A. E., 2000, PASP, 112, 1383
Eggenberger A., Udry S., Mayor M., 2004, A&A, 417, 353
Fabrycky D., Tremaine S., 2007, ApJ, 669, 1298
Fortney J. J., Marley M. S., Barnes J. W., 2007, ApJ, 659,
1661
Fried D. L., 1978, Journal of the Optical Society of America
(1917-1983), 68, 1651
Fruchter A. S., Hook R. N., 2002, PASP, 114, 144
Gillon M., Lanotte A. A., Barman T., Miller N., Demory
B.-O., Deleuil M., et al., 2010, A&A, 511, A3
Guillot T., 2005, Annual Review of Earth and Planetary
Sciences, 33, 493
Hardy J. W., 1998, Adaptive Optics for Astronomical Tele-
scopes
H´ebrard G., D´esert J.-M., D´ıaz R. F., et al., 2010, A&A,
516, A95+
Hinkley S., Oppenheimer B. R., Soummer R., Sivaramakr-
ishnan A., Roberts Jr. L. C., Kuhn J., Makidon R. B.,
Perrin M. D., Lloyd J. P., Kratter K., Brenner D., 2007,
ApJ, 654, 633
Høg E., Fabricius C., Makarov V. V., Urban S., Corbin
T., Wycoff G., Bastian U., Schwekendiek P., Wicenec A.,
2000, A&A, 355, L27
Ida S., Lin D. N. C., 2004, ApJ, 604, 388
Juri´c M., Ivezi´c Z., Brooks A., et al., 2008, ApJ, 673, 864
Kraus A. L., Hillenbrand L. A., 2007, AJ, 134, 2340
Law N. M., Mackay C. D., Baldwin J. E., 2006, A&A, 446,
Baldwin J. E., Tubbs R. N., Cox G. C., Mackay C. D.,
739
Wilson R. W., Andersen M. I., 2001, A&A, 368, L1
Baraffe I., Chabrier G., Allard F., Hauschildt P. H., 1998,
Lee M. H., Peale S. J., 2002, ApJ, 567, 596
Lin D. N. C., Bodenheimer P., Richardson D. C., 1996,
A&A, 337, 403
Nature, 380, 606
Bergfors C., Brandner W., Daemgen S., Biller B., Hippler
S., Janson M., Kudryavtseva N., Geissler K., Henning T.,
Kohler R., 2013, MNRAS, 428, 182
Boccaletti A., Chauvin G., Lagrange A.-M., Marchis F.,
2003, A&A, 410, 283
Boccaletti A., Riaud P., Baudoz P., Baudrand J., Rouan
D., Gratadour D., Lacombe F., Lagrange A.-M., 2004,
PASP, 116, 1061
Liu X., Burrows A., Ibgui L., 2008, ApJ, 687, 1191
Mackay C. D., Baldwin J., Law N., Warner P., 2004, in
A. F. M. Moorwood & M. Iye ed., Society of Photo-
Optical Instrumentation Engineers (SPIE) Conference Se-
ries Vol. 5492 of Society of Photo-Optical Instrumenta-
tion Engineers (SPIE) Conference Series, High-resolution
imaging in the visible from the ground without adaptive
optics: new techniques and results. pp 128 -- 135
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Lucky Imaging of transiting planet hosts
13
Madhusudhan N., Harrington J., Stevenson K. B.,
Nymeyer S., Campo C. J., Wheatley P. J., et al., 2011,
Nature, 469, 64
Majewski S. R., Skrutskie M. F., Weinberg M. D., Os-
theimer J. C., 2003, ApJ, 599, 1082
Marois C., Nadeau D., Doyon R., Racine R., Walker
G. A. H., 2003, in Mart´ın E., ed., Brown Dwarfs Vol. 211 of
IAU Symposium, Differential Simultaneous Imaging and
Faint Companions: TRIDENT First Results from CFHT.
p. 275
Marzari F., Nelson A. F., 2009, ApJ, 705, 1575
McLaughlin D. B., 1924, ApJ, 60, 22
Morton T. D., Johnson J. A., 2011, ApJ, 729, 138
Nagasawa M., Ida S., Bessho T., 2008, ApJ, 678, 498
Narita N., Kudo T., Bergfors C., Nagasawa M., et al., 2010,
PASJ, 62, 779
Narita N., Takahashi Y. H., Kuzuhara M., Hirano T., Sue-
naga T., Kandori R., Kudo T., et al., 2012, PASJ, 64,
L7
Pont F., H´ebrard G., Irwin J. M. o., 2009, A&A, 502, 695
Rasio F. A., Ford E. B., 1996, Science, 274, 954
Rossiter R. A., 1924, ApJ, 60, 15
Schlaufman K. C., 2010, ApJ, 719, 602
Schroter S., Czesla S., Wolter U., Muller H. M., Huber
K. F., Schmitt J. H. M. M., 2011, A&A, 532, A3
Staley T. D., Mackay C. D., 2010, in Ground-based and
Airborne Instrumentation for Astronomy III No. 7735 in
Proc. SPIE, Data reduction strategies for lucky imaging
Stetson P. B., 1987, PASP, 99, 191
Takeda G., Kita R., Rasio F. A., 2008, ApJ, 683, 1063
Tokovinin A., 2002, PASP, 114, 1156
Triaud A. H. M. J., Collier Cameron A., Queloz D., An-
derson D. R., Gillon M., Hebb L., Hellier C., Loeillet
B., Maxted P. F. L., Mayor M., Pepe F., Pollacco D.,
S´egransan D., Smalley B., Udry S., West R. G., Wheatley
P. J., 2010, A&A, 524, A25+
Tubbs R. N., Baldwin J. E., Mackay C. D., Cox G. C.,
2002, A&A, 387, L21
Watson C. A., Littlefair S. P., Diamond C., Collier
Cameron A., Fitzsimmons A., Simpson E., Moulds V.,
Pollacco D., 2011, MNRAS, 413, L71
Weidenschilling S. J., Marzari F., 1996, Nature, 384, 619
Winn J. N., 2010, Exoplanet Transits and Occultations. pp
55 -- 77
Winn J. N., Fabrycky D., Albrecht S., Johnson J. A., 2010,
ApJ, 718, L145
Winn J. N., Howard A. W., Johnson J. A., et al., 2009,
ApJ, 703, 2091
Winn J. N., Howard A. W., Johnson J. A., Marcy G. W.,
Isaacson H., Shporer A., Bakos G. ´A., Hartman J. D.,
Holman M. J., Albrecht S., Crepp J. R., Morton T. D.,
2011, AJ, 141, 63
Winn J. N., Johnson J. A., Albrecht S., Howard A. W.,
Marcy G. W., Crossfield I. J., Holman M. J., 2009, ApJ,
703, L99
Wu Y., Murray N., 2003, ApJ, 589, 605
Zacharias N., Monet D. G., Levine S. E., Urban S. E.,
Gaume R., Wycoff G. L., 2005, VizieR Online Data Cat-
alog, 1297, 0
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
1702.06356 | 1 | 1702 | 2017-02-21T12:38:22 | Simulating the Smallest Ring World of Chariklo | [
"astro-ph.EP"
] | A ring system consisting of two dense narrow rings has been discovered around Centaur Chariklo. The existence of these rings around a small object poses various questions, such as their origin, stability, and lifetime. In order to understand the nature of Chariklo's rings, we perform global $N$-body simulations of the self-gravitating collisional particle rings for the first time. We find that Chariklo should be denser than the ring material to avoid the rapid diffusion of the rings. If Chariklo is denser than the ring material, fine spiral structures called self-gravity wakes occur in the inner ring. These wakes accelerate the viscous spreading of the ring significantly and they typically occur on timescales of about $100\,\mathrm{years}$ for m-sized ring particles, which is considerably shorter than the timescales suggested in previous studies. The existence of these narrow rings implies smaller ring particles or the existence of shepherding satellites. | astro-ph.EP | astro-ph |
Simulating the Smallest Ring World of Chariklo
Shugo Michikoshi, and Eiichiro Kokubo
[email protected], and [email protected]
ABSTRACT
A ring system consisting of two dense narrow rings has been discovered around
Centaur Chariklo. The existence of these rings around a small object poses various
questions, such as their origin, stability, and lifetime. In order to understand the na-
ture of Chariklo's rings, we perform global N -body simulations of the self-gravitating
collisional particle rings for the first time. We find that Chariklo should be denser than
the ring material to avoid the rapid diffusion of the rings. If Chariklo is denser than
the ring material, fine spiral structures called self-gravity wakes occur in the inner ring.
These wakes accelerate the viscous spreading of the ring significantly and they typically
occur on timescales of about 100 years for m-sized ring particles, which is considerably
shorter than the timescales suggested in previous studies. The existence of these narrow
rings implies smaller ring particles or the existence of shepherding satellites.
1.
Introduction
Recently, two dense narrow rings around Centaur Chariklo were discovered by occultation
observation (Braga-Ribas et al. 2014). A ring around Centaur Chiron was also proposed (Ruprecht
et al. 2015; Ortiz et al. 2015). These observations suggest that rings around large Centaurs may
not be as rare as previously thought.
Several mechanisms for the formation of Chariklo's rings have been proposed: collisional ejec-
tion from the parent body, satellite disruption, and out gassing (Pan & Wu 2016). The tidal
disruption of a differentiated object during a close encounter with a giant planet can also result in
the formation of a ring (Hyodo et al. 2016). From the numerical integration of the Chariklo orbit
it was suggested that the possibility of a close encounter with a giant planet that is able to disrupt
Chariklo's rings is low (Araujo et al. 2016). The formation mechanism of the rings is still not well
understood. In order to understand the origin of the ring system, we need to investigate the ring
structure in detail.
1 Center for Computational Sciences, University of Tsukuba, Tsukuba, Ibaraki 305-8577, Japan
2 Division of Theoretical Astronomy, National Astronomical Observatory of Japan, Osawa, Mitaka, Tokyo 181-
8588, Japan
– 2 –
Observations have found a difference in the inner ring widths between ingress and egress
(Braga-Ribas et al. 2014). The width variation indicates that the inner ring has a finite eccentricity.
On the other hand, Chariklo has large oblateness, which should lead to the rapid differential
precession of the ring, resulting in a circular ring. Thus, there must be some mechanism keeping
the ring eccentricity. To explain the eccentric ring around Uranus, apse alignment due to ring
self-gravity has been proposed (Goldreich & Tremaine 1979a,b). By applying the same model to
Chariklo's rings, the ring mass and particle size have been estimated (Pan & Wu 2016). From
these estimates, they found that the inner ring is slightly gravitationally unstable in the context of
Toomre stability (Toomre 1964).
Local N -body simulations of Saturn's rings revealed that self-gravity wakes exist in Saturn's
A and B rings (e.g., Salo 1992, 1995; Michikoshi & Kokubo 2011; Michikoshi et al. 2015). The
self-gravity wakes are non-axisymmetric small spiral structures formed by gravitational instability,
which enable efficient angular momentum transfer. Thus, it is important to elucidate whether
self-gravity wakes exist in Chariklo's rings. The optical depth in Chariklo's inner ring is about
0.38, which is marginal for the formation of self-gravity wakes (Salo 1995). The critical optical
depth for the formation of self-gravity wakes depends on the distance from the central body and
the properties of the ring particles.
In order to investigate Chariklo's ring system, we perform
global N -body simulations of the two narrow rings with full self-gravity for the first time. Section
2 describes the model and the simulation method, Section 3 presents the results of the numerical
simulations, and Section 4 contains a summary and a discussion.
2. Simulation Method
Chariklo has a large oblateness of about 0.2 (Braga-Ribas et al. 2014), resulting in a large
gravitational moment J2, which causes the differential precession of the rings. We here focus on the
formation of structures over the dynamical timescale. Thus, we neglect the oblateness. The latest
observations and analysis show that Chariklo has a shape with volume equivalent to a sphere of
radius 125 km and its density is 0.8 or 1.2 g cm−3 (Leiva et al. 2016). We adopt a spherical Chariklo
with radius 125 km and density ρC = 1.0 g cm−3. Chariklo has the inner dense ring, with different
optical depths and radial widths for the ingress (W = 6.16 km, τ = 0.449) and egress (W = 7.17 km,
τ = 0.317). This indicates that the inner ring may have a finite eccentricity. However, in this paper,
we assume that the inner ring is circular and uniform with W = 6.7 km and τ = 0.38 as a first step.
Similarly, we assume that the outer ring is circular and uniform with W = 3.5 km and τ = 0.06.
The inner and outer rings are located at distances of a = 390.6 km and 404.8 km from the center
of Chariklo, respectively.
The properties of ring particles are not well constrained. For the sake of simplicity, we assume
that all particles have the same mass mp and radius rp. In Saturn's rings, the typical particle size
varies between a centimeter and a few meters (French & Nicholson 2000). Furthermore, considering
the apse alignment of the ring by its self-gravity, Pan & Wu (2016) estimated the ring mass,
– 3 –
which corresponds to a few meter-sized particles. We assume rp = 2.5–10 m, and density ratios
ρp/ρC = 0.05, 0.10, 0.25, 0.50, 0.75, and 1.00. The number of ring particles is 21 to 345 million. The
basic ring dynamics is controlled by the two non-dimensional ring parameters τ and rH, where rH
is the scaled Hill radius of ring particles given by rH = rH/2rp with rH = (2mp/3MC)1/3a (Salo
1995), where MC is Chariklo's mass. Note that rH is independent of the particle size.
The equation of motion of particle i is
ri3 −(cid:88)
ri
j(cid:54)=i
d2ri
dt2 = −GMC
Gmp
rij
ij + 2
(r2
g)3/2
+ fcol,
(1)
where ri is the position vector of particle i, rij = ri − rj, rij = rij, g is the softening length,
and fcol is the collisional force. We adopt the soft sphere model as the collision model (Salo 1995).
In this model, a collision is described as a damped oscillation. This model has a free parameter,
the collision duration tcol, which should be sufficiently shorter than the dynamical timescale. The
collision duration is tcol = 0.0025tK, where tK is the orbital period. The restitution coefficient is
= 0.1. We set g = 0.1rp.
We integrate the equation of motion with the leapfrog method. To reduce the computational
time, as in Salo (1995), we consider two different timesteps, ∆tgrav and ∆tcol. The timestep ∆tcol
is for collisions, and we integrate the equation of motion with ∆tcol. We update the gravitational
acceleration with ∆tgrav, which is set to be longer than ∆tcol. We adopt ∆tgrav = 0.005tK, which is
sufficiently small for resolving gravitational interactions among particles, and ∆tcol = ∆tgrav/32 =
tcol/16. We use the tree algorithm for the gravitational force calculation (Barnes & Hut 1986).
We adopt the opening angle θ = 0.5, and employ the N -body simulation library, Framework for
Developing Particle Simulator (FDPS) (Iwasawa et al. 2016) with the Phantom-GRAPE module
(Tanikawa et al. 2012, 2013).
3. Results
3.1. Formation of Self-gravity Wakes
Figure 1 shows the ring at t = 10tK, where ρp/ρC = 0.50 and rp = 5 m; no large-scale
structures are visible. For dynamical timescales, the ring keeps its original global shape. However,
we find small-scale structures, namely, self-gravity wakes, in the inner ring (Fig. 2). Initially, the
ring particles are distributed uniformly. At t = 1tK, fluctuations in the ring surface density grow
due to gravitational instability, and self-gravity wakes form at t = 2tK. After t = 3tK, the self-
gravity wakes are formed and destroyed continuously. The self-gravity and collision of particles form
particle aggregates, while the differential rotation tears them apart. These competing processes are
the cause of self-gravity wakes.
The spatial scale of self-gravity wakes is characterized by the critical wavelength of the gravi-
– 4 –
tational instability (Toomre 1964; Salo 1995; Daisaka & Ida 1999)
λcr =
4π2GΣ
Ω2 = 0.36
ai
(cid:16)
(cid:17)3(cid:16) rC
(cid:17)−3(cid:18) ρp/ρC
(cid:19)(cid:16) τ
(cid:17)(cid:16) rp
(cid:17)
390.6 km
125 km
0.5
0.38
5 m
km,
(2)
where Σ is the ring surface density and Ω = 2π/tK is the Kepler angular frequency.
On the other hand, in the outer ring, the self-gravity wakes do not develop. Since the optical
depth in the outer ring is small, the energy dissipation by collisions is insufficient. Thus, the random
velocity of ring particles is high and the ring is gravitationally stable. Since the outer ring is near
the Roche limit, small aggregates are visible. The tidal force is comparable with their self-gravity.
Thus they may be temporal.
3.2. Ring Particle Properties
We constrain the properties of ring particles with those of the self-gravity wakes. Figure 3 shows
the structure of the rings at t = 10 tK for the models where ρp/ρC = 0.05, 0.10, 0.25, 0.50, 0.75 and
1.00 with rp = 5 m. In the ρp/ρC = 0.75, and 1.00 models, large particle aggregates form and the
ring spreads rapidly over the dynamical timescale. Thus, this model may not correspond to the real
ring. The scaled Hill radius determines whether a particle pair can gravitationally bind (Ohtsuki
1993; Salo 1995; Karjalainen & Salo 2004), which is given as
(cid:16)
a
(cid:17)(cid:16) rC
(cid:17)−1(cid:18) ρp
(cid:19)1/3
rH = 1.36
(3)
Studies of Saturn's rings showed that gravitationally bound aggregates form if rH (cid:38) 1.1 (Ohtsuki
1993; Salo 1995; Karjalainen & Salo 2004). Therefore, the necessary condition for ring formation
is ρp/ρC < 0.52, in other words, the particle density should be lower than half of that of Chariklo.
390.6 km
125 km
.
ρC
In the inner ring, self-gravity wakes appear for ρp/ρC = 0.25 and 0.5, while they do not appear
for ρp/ρC = 0.05 and 0.1. As the particle density increases, the spatial scale of the self-gravity wakes
increases, as shown in Equation (2). For τ = 0.38, the self-gravity wakes form under the condition
0.65 (cid:46) rH (cid:46) 1.1 (Salo 1995; Ohtsuki & Emori 2000; Daisaka et al. 2001), which is independent of
the particle radius. If rH (cid:46) 0.65, the self-gravity wake formation is suppressed, which corresponds
to ρp/ρC < 0.11. If Chariklo has the typical Centaur density, ρC (cid:39) 1.0 g cm−3, the particle density
needs to be less than 0.1 g cm−3 for the suppression of self-gravity wakes.
Figure 4 shows the dependence of the spatial scale of self-gravity wakes on rp with ρp/ρC = 0.5,
rp = 2.5, 5, 7.5, and 10 m. In all the models, the self-gravity wakes develop, which is consistent with
the condition for self-gravity wake formation discussed above. As indicated by Equation (2), the
spatial scale increases with rp. The ring width for larger particles is wider than that with smaller
particles. This is because the viscosity due to self-gravity wakes increases with wake size.
The Chariklo ring was discovered by stellar occultation (Braga-Ribas et al. 2014). The pro-
jected star radius for Chariklo was estimated at around 1 km. Figure 5 shows a simulation of the
– 5 –
Fig. 1.- Snapshots of the simulated ring in the x–y plane at t = 10tK, where the particle density
and radius are ρp/ρC = 0.5 and rp = 5 m, respectively. The left panel (a) shows the overall structure
of the ring, while the panels (b), (c), (d) show enlarged views of ring sections.
– 6 –
(a) t = 0tK
(b) t = 1tK
(c) t = 2tK
(d) t = 3tK
Fig. 2.- Snapshots of the inner ring at (a) t = 0, (b) 1tK, (c) 2tK, and (d) 3tK for the model with
ρp/ρC = 0.5 and rp = 5 m.
– 7 –
(a) ρp/ρC = 0.05
(b) ρp/ρC = 0.10
(c) ρp/ρC = 0.25
(d) ρp/ρC = 0.50
(e) ρp/ρC = 0.75
(f) ρp/ρC = 1.00
Fig. 3.- Snapshots of the inner ring at t = 10tK for the models with (a) ρp/ρC = 0.05, (b) 0.1,
(c) 0.25, (d) 0.50, (e) 0.75, and (f) 1.0 with rp = 5 m.
– 8 –
(a) rp = 2.5 m
(b) rp = 5 m
(c) rp = 7.5 m
(d) rp = 10 m
Fig. 4.- Snapshots of the inner ring at t = 10tK for the models with (a) rp = 2.5 m, (b) 5 m, (c)
7.5 m, and (d) 10 m with ρp/ρC = 0.50.
– 9 –
dynamical optical depth τ = 4Σ/3rpρp with spatial resolutions with radii ∆r = 0.25, 0.50, and
1.0 km for the model where ρp/ρC = 0.5 and rp = 10 m. The critical wavelength is λcr = 0.72 km.
We can observe fluctuations due to self-gravity wakes in the ∆r = 0.25 km and 0.5 km models, while
the optical depth is almost smooth in the ∆r = 1.0 km model. If the spatial scale of self-gravity
wakes is larger than the star projected radius, the fluctuation due to self-gravity wakes may be
detected by occultation observations. However, the observations show the smooth distribution.
Thus, the spatial scale of self-gravity wakes would be smaller than 1 km, which leads to
(cid:18) ρp/ρC
(cid:19)−1(cid:16) τ
(cid:17)−1
0.5
0.38
rp (cid:46) 13.7
m,
(4)
which is consistent with the estimate by Pan & Wu (2016). In future observations, if we detect
spatial variation of the optical depth, this will confirm the existence of self-gravity wakes and give
a lower boundary for the particle size.
If we consider the ring mass estimated by Pan & Wu (2016), we can give the lower boundary
of the particle size. The surface density is estimated as few × 100 g cm−2 from the apse alignment
argument. From the ring formation criterion ρp/ρC < 0.52, we give the following constraint
(cid:18)
(cid:19)
rp (cid:38) 3.8
Σ
100 g cm−2
m.
(5)
3.3. Ring Lifetime
The self-gravity wakes cause efficient radial diffusion in the inner ring. The effective viscosity
of self-gravity wakes was obtained by N -body simulations (Daisaka et al. 2001; Tanaka et al. 2003),
which is given as νsgw (cid:39) 26r5
HG2Σ2/Ω3. Using this, we obtain the diffusion time of the inner ring
as
(cid:19)−11/3
years,
(6)
(cid:17)−2(cid:18)
(cid:16) rp
1 m
tinner ≡ W 2
ν
(cid:39) 113
ρp
0.5 g cm−3
where we assume W = 6.7 km, τ = 0.38, and a = 390.6 km. This diffusion time is much shorter
than in previous studies (Braga-Ribas et al. 2014; Pan & Wu 2016) where the self-gravity wakes
are not taken into account.
For the outer ring we estimate the diffusion time using the collisional viscosity, νcol (cid:39) r2
pΩτ ,
which is
touter (cid:39) 7.1 × 104(cid:16) rp
(cid:17)−2
1 m
years,
(7)
where we assume W = 3.5 km, τ = 0.06, and a = 404.8 km. This is roughly consistent with the
results of previous studies (Braga-Ribas et al. 2014; Pan & Wu 2016).
Equation (6) shows that the lifetime of the inner ring is much shorter than the typical dynamical
lifetime of Centaurs, ∼ 106 years, if the particle size is on the order of 1 m. If the rings were formed
– 10 –
Fig. 5.- Simulation of the dynamical optical depth along the line y = 0 at t = 10 tK for the model
where ρp/ρC = 0.5 and rp = 10 m. The spatial resolutions are ∆r = 0.25 km (dotted), ∆r = 0.5 km
(dashed), and 1.0 km (solid).
00.10.20.30.40.50.60.70.80.9385390395400405410OpticalDepthRadius[km]∆r=0.25km0.5km1.0km– 11 –
at the same time when Chariklo was scattered into the Centaur region, this significantly short ring
lifetime would be inconsistent. One possible solution to this contradiction is that Chariklo's rings
are very young. Another solution is that the ring consists of small particles, such as particles of
less than 1 cm. It is also possible that the narrow ring is due to shepherding satellites, which will
be discussed below (Braga-Ribas et al. 2014).
4. Summary and Discussion
We performed global N -body simulations of Chariklo's rings and investigated their structure.
We found that in order for Chariklo to host rings instead of particle aggregates its density should be
larger than that of particles. Under this condition, the self-gravity wakes inevitably develop in the
inner ring independently of the particle size, while their spatial scale depends on the particle size.
For m-sized ring particles, the timescale of ring viscous spreading due to the self-gravity wakes is on
the order of 100 years, which is much shorter than that estimated in previous studies (Braga-Ribas
et al. 2014; Pan & Wu 2016).
Our simulations predict that Chariklo is denser than the ring particles, and the ring particles
In Saturn's rings, from comparisons of
are also less dense than ice, similar to Saturn's rings.
observations and N -body simulations, a particle density less than that of ice has been proposed
(Salo et al. 2004; Michikoshi et al. 2015). The higher density of Chariklo than that of the ring
particles suggests that Chariklo may have a dense core. Previous studies have hypothesized that
the ring material originates from the stripped icy mantle of a differentiated body (Hyodo et al.
2016), which is consistent with our findings.
The timescale of viscous ring spreading (Equation (6)) suggests three possibilities: a very
young ring (∼ 1–100 years), smaller ring particles, or existence of shepherding satellites.
If the
ring is young, because a close encounter with a giant planet is rare (Araujo et al. 2016), the ring
formation by the tidal interaction with giant planets may be difficult. Then the other mechanisms,
such as out gassing, would be preferable (Pan & Wu 2016). If the ring particles are smaller than ∼ 1
cm, the inner ring can last longer than 106 years. However, this small particle size is inconsistent
with the estimate from the apse alignment of the ring (Pan & Wu 2016). For the m-sized particles
the existence of shepherding satellites is required to counteract the viscous ring spreading. The
minimum mass of the shepherding satellite depends on the distance from the inner ring edge d and
the particle size and density (Goldreich & Tremaine 1982),
(cid:18) νd3
(cid:19)1/2
Ωa5
Ms (cid:39) 4.1
MC = 1.3 × 1017
(cid:18) d
(cid:19)3/2(cid:18)
100 km
(cid:19)11/6(cid:16) rp
(cid:17)
1 m
ρp
0.5 g cm−3
where we assumed the viscosity due to the self-gravity wakes, which is ν ∝ Σ2. The size of a satellite
with mass 1017g is on the order of kilometers. Furthermore, the shepherding satellite hypothesis
may be preferable because it could also explain the ring eccentricity.
g,
(8)
– 12 –
The particle size is a key parameter for determining the dynamical property of Chariklo's
rings. However, the particle size has not yet been constrained observationally. In future, if the
self-gravity wakes are detected by higher resolution occultation observations, they will be able to
provide a lower limit for the particle size.
Our simulations suggest that the formation of self-gravity wakes is a general process in dense
narrow rings. For example, the Huygens ringlet in the Cassini division of Saturn's ring and the
ring of Uranus have a sufficiently high optical depth to form self-gravity wakes. However, their
spatial scale is far below the observational limit today.
In this study, we assumed the circular ring, though the inner ring may have a finite eccentricity.
The effect of the eccentricity on the self-gravity wake dynamics has not been examined. In the
subsequent work, we plan to investigate its effect with more realistic ring models considering, for
example, the size distribution of particles, the restitution coefficient that depends on the collisional
velocity (Bridges et al. 1984), and shepherding satellites.
Numerical computations were carried out on ATERUI (Cray XC30) at the Center for Compu-
tational Astrophysics, National Astronomical Observatory of Japan.
REFERENCES
Araujo, R. A. N., Sfair, R., & Winter, O. C. 2016, ApJ, 824, 80
Barnes, J. & Hut, P. 1986, Nature, 324, 446
Braga-Ribas, F., Sicardy, B., Ortiz, J. L., Snodgrass, C., Roques, F., Vieira-Martins, R., Camargo,
J. I. B., Assafin, M., Duffard, R., Jehin, E., Pollock, J., Leiva, R., Emilio, M., Machado,
D. I., Colazo, C., Lellouch, E., Skottfelt, J., Gillon, M., Ligier, N., Maquet, L., Benedetti-
Rossi, G., Gomes, A. R., Kervella, P., Monteiro, H., Sfair, R., El Moutamid, M., Tancredi,
G., Spagnotto, J., Maury, A., Morales, N., Gil-Hutton, R., Roland, S., Ceretta, A., Gu, S.-
H., Wang, X.-B., Harpsøe, K., Rabus, M., Manfroid, J., Opitom, C., Vanzi, L., Mehret, L.,
Lorenzini, L., Schneiter, E. M., Melia, R., Lecacheux, J., Colas, F., Vachier, F., Widemann,
T., Almenares, L., Sandness, R. G., Char, F., Perez, V., Lemos, P., Martinez, N., Jørgensen,
U. G., Dominik, M., Roig, F., Reichart, D. E., Lacluyze, A. P., Haislip, J. B., Ivarsen, K. M.,
Moore, J. P., Frank, N. R., & Lambas, D. G. 2014, Nature, 508, 72
Bridges, F. G., Hatzes, A., & Lin, D. N. C. 1984, Nature, 309, 333
Daisaka, H. & Ida, S. 1999, Earth, Planets, and Space, 51, 1195
Daisaka, H., Tanaka, H., & Ida, S. 2001, Icarus, 154, 296
French, R. G. & Nicholson, P. D. 2000, Icarus, 145, 502
Goldreich, P. & Tremaine, S. 1979a, AJ, 84, 1638
– 13 –
-. 1979b, Nature, 277, 97
-. 1982, ARA&A, 20, 249
Hyodo, R., Charnoz, S., Genda, H., & Ohtsuki, K. 2016, ApJ, 828, L8
Iwasawa, M., Tanikawa, A., Hosono, N., Nitadori, K., Muranushi, T., & Makino, J. 2016, PASJ,
68, 54
Karjalainen, R. & Salo, H. 2004, Icarus, 172, 328
Leiva, R., Sicardy, B., Berard, D., Meza Quispe, E., camargo, j., Assafin, M., Braga-Ribas, F.,
Vieira-Martins, R., Maquet, L., Colas, F., Sickafoose, A. A., Bath, K.-L., & Dauvergne, J.-
L. 2016, in AAS/Division for Planetary Sciences Meeting Abstracts, Vol. 48, AAS/Division
for Planetary Sciences Meeting Abstracts, 203.07
Michikoshi, S., Fujii, A., Kokubo, E., & Salo, H. 2015, ApJ, 812, 151
Michikoshi, S. & Kokubo, E. 2011, ApJ, 732, L23+
Ohtsuki, K. 1993, Icarus, 106, 228
Ohtsuki, K. & Emori, H. 2000, AJ, 119, 403
Ortiz, J. L., Duffard, R., Pinilla-Alonso, N., Alvarez-Candal, A., Santos-Sanz, P., Morales, N.,
Fern´andez-Valenzuela, E., Licandro, J., Campo Bagatin, A., & Thirouin, A. 2015, A&A,
576, A18
Pan, M. & Wu, Y. 2016, ApJ, 821, 18
Ruprecht, J. D., Bosh, A. S., Person, M. J., Bianco, F. B., Fulton, B. J., Gulbis, A. A. S., Bus,
S. J., & Zangari, A. M. 2015, Icarus, 252, 271
Salo, H. 1992, Nature, 359, 619
-. 1995, Icarus, 117, 287
Salo, H., Karjalainen, R., & French, R. G. 2004, Icarus, 170, 70
Tanaka, H., Ohtsuki, K., & Daisaka, H. 2003, Icarus, 161, 144
Tanikawa, A., Yoshikawa, K., Nitadori, K., & Okamoto, T. 2013, New A, 19, 74
Tanikawa, A., Yoshikawa, K., Okamoto, T., & Nitadori, K. 2012, New A, 17, 82
Toomre, A. 1964, ApJ, 139, 1217
This preprint was prepared with the AAS LATEX macros v5.2.
|
1210.6218 | 2 | 1210 | 2012-10-24T20:23:04 | A preliminary study of the linear relationship between monthly averaged daily solar radiation and daily thermal amplitude in the north of Buenos Aires province | [
"astro-ph.EP"
] | Using irradiance and temperature measurements obtained at the Facultad Regional San Nicol\'as of UTN, we performed a preliminary study of the linear relationship between monthly averaged daily solar radiation and daily thermal amplitude. The results show a very satisfactory adjustment (R = 0.848, RMS = 0.066, RMS% = 9.690 %), even taking into account the limited number of months (36). Thus, we have a formula of predictive nature, capable of estimating mean monthly solar radiation for various applications. We expect to have new data sets to expand and improve the statistical significance of these results. | astro-ph.EP | astro-ph | A PRELIMINARY STUDY OF THE LINEAR RELATIONSHIP BETWEEN MONTHLY
AVERAGED DAILY SOLAR RADIATION AND DAILY THERMAL AMPLITUDE IN
THE NORTH OF BUENOS AIRES PROVINCE
RELACI ÓN LINEAL ENTRE LOS PROMEDIOS MENSUALES DE RA DIACI ÓN SOLAR
Y AMPLITUD TÉRMICA EN SAN NICOLÁS,
NORTE DE LA PROVINCIA DE BUENOS AIRES
R. Cionco, N. Quaranta1, R. Rodriguez
Universidad Tecnológica Nacional-Facultad Regional San Nicolás, Colón 332, San Nicolás, c.p. 2900, Buenos
Aires
Tel.: +54 -336- 4420830; fax: +54- 0336 4425266 e-mail: [email protected]
1 Investigador CIC-PBA
ABSTRACT
Using irradiance and temperature measurements obtained at the Facultad Regional San Nicolás of UTN, we
performed a preliminary study of the linear relationship between monthly averaged daily solar radiation and
daily thermal amplitude. The results show a very satisfactory adjustment (R = 0.848, RMS = 0.066, RMS% =
9.690 %), even taking into account the limited number of months (36). Thus, we have a formula of predictive
nature, capable of estimating mean monthly solar radiation for various applications. We expect to have new data
sets to expand and improve the statistical significance of these results.
Keywords: daily solar energy, prediction, thermal amplitude, north of Buenos Aires province.
RESUMEN: Utilizando medidas de radiación solar global y temp eratura obtenidas en la Facultad Regional San
Nicolás de UTN, se realiz ó un estudio preliminar de la relación lineal entre los promedios mensuales d e
radiación solar diaria y las medias mensuales de am plitud térmica diaria. Los resultados muestran un ajuste muy
satisfactorio (R = 0,848; RMS = 0,066; RMS% = 9,690 %), aun teniendo en cuenta el número limitado de meses
(36). De esta forma, se dispone de una fórmula de c arácter predictivo, capaz de estimar medias mensuales de
radiación solar para aplicaciones diversas. Se espe ra disponer de nuevas series de datos para ampliar y mejorar la
significación estad ística de estos resultados.
Palabras clave: radiación solar diaria, predicción, amplitud térmic a, norte de Buenos Aires.
INTRODUCCIÓN
La Universidad Tecnológica Nacional a través del Gr upo de Estudios Ambientales de la Facultad Regional San
Nicolás, realiza desde hace algunos años un trabajo de monitoreo de calidad de aire mediante convenios
espec íficos con instituciones públicas y privadas; en particular, con las principales empresas del importante
cintur ón industrial de la zona. Como parte de esta actividad, se obtienen datos meteorológicos con div ersas
estaciones automatizadas, una de las cuales, ha realizado mediciones de irradiancia solar. Esos datos han
comenzado a ser liberados para usos diversos, lo cual motiv ó el presente trabajo.
El norte de la provincia de Buenos Aires, especialmente la zona de San Nicolás de los Arroyos (j = - 33o 19´, l
= - 60o 12´) es una de las regiones de mayor actividad ind ustrial y agr ícola del pa ís, además de presentar gr an
variabilidad climática asociada a fenómenos como El Niño (Rusticucci y Vargas, 2002). La medición dire cta de
la radiación solar en la zona y en la República Arg entina en general, es escasa (Podestá et. al, 2004; Raichijk et.
al, 2005); por lo tanto, el seguimiento y estudio de este parámetro es importante especialmente para aplicaciones
agr ícolas, como por ejemplo dinámica de crecimiento de cultivos (Meinke et al., 1995), de contaminació n
atmosférica (dispersión de PM10, reacciones formado ras de NOx, SOx y formación de ozono; Seinfeld, 200 1),
además de tener interés primario en la cuantificación y seguimiento del recurso solar per se. No es de
conocimiento de los autores que existan trabajos publicados de cuantificación directa de la radiación
solar para la
zona. Podestá et. al (2004), presentan resultados de la determinación de la radiación solar diaria a p
artir de las
medidas de energía solar global de las estaciones d el INTA-SMN, obtenidas durante los años 1981 y 1982 en
Pergamino, a unos 65 km al sudoeste de San Nicolás. Sin embargo, existen otros trabajos relevantes que
involucran a la zona a la hora de cuantificar el recurso solar y de determinar fórmulas predictivas, e stos son los
de Righini y Grossi Gallegos (2003), Raichijk et al. (2005) y Raichijk y Grossi Gallegos (2007). Aunque los
datos disponibles para este análisis pertenecen solamente al trienio 2006-2008, la posible discontinuidad en estas
mediciones o la falta de disponibilidad de datos a futuro motivó la realización de un trabajo prelimin ar,
utilizando los datos existentes para los años mencionados. No se dispone hasta el momento de datos más
actuales.
La necesidad de conocer valores de radiación solar global y las dificultades generales para obtenerlos, han
promovido la búsqueda de relaciones que permitan realizar predicciones a partir de determinados parámetros
meteorológicos. La más común para la predicción de estos valores, es la bien conocida fórmula de Ångst
r öm
(Ångstr öm, 1924) posteriormente modificada por Pres cott (Prescott, 1940), que relaciona la radiación s olar
diaria Q de un sitio, generalmente en unidades del valor a tope de atmósfera o extraterrestre ( Qt); o sea el
cociente Q/Qt, el cual, por brindar una cuantificación de la radi ación solar medida en tierra con respecto a la que
te óricamente llega a un plano horizontal (respecto a la vertical local) por sobre la atmósfera, suele denominarse
transmisividad atmosférica o coeficiente de claridad K. En el modelo predictivo de Ångstr öm-Prescott (que
originalmente fue elaborado para relacionar promedios mensuales de K, pero actualmente es usado en diversas
escalas de tiempo; ver por ejemplo, Podestá et al., 2004), la variable independiente es la heliofanía relativa, que
generalmente es un parámetro tanto o más escaso que los datos de radiación solar diaria. Por lo tanto,
con la
finalidad de poseer fórmulas más accesibles que permitan predecir valores de Q, se han ensayado otros métodos
que tienen en cuenta variables meteorológicas de ma yor disponibilidad o más fácilmente obtenibles.
Particularmente, los métodos emp íricos basados en la relaci ón entre K y la amplitud térmica diaria (la diferencia
entre las temperaturas máxima y mínima del aire u otras combinaciones entre ellas), resultan muy convenientes,
por ser la amplitud térmica una de las variables meteorológicas más ampliamente medidas (ver por ejemp lo,
Meinke et al., 1995; Meza y Varas, 2000; Raichijk et al., 2005; Bandyopadhyay et al., 2008).
El presente trabajo está básicamente inspirado en el de Raichijk et al. (2005), quienes han relacionado valores
mensuales promedio de Q en distintas localidades de Argentina, mediante un análisis de regresión lineal (método
de Hargreaves o Hargreaves-Samani; citados por Bandyopadhyay et al., 2008) con los promedios mensuales de
amplitud térmica diaria, de la siguiente forma:
<Q>/<Qt> = a + b <DT>1/2 (1)
donde a y b son constantes reales; <Q> es el promedio mensual de los valores calculados de Q; <Qt> es el
promedio mensual de los valores te óricos Qt y <DT>1/2 es la ra íz cuadrada del promedio mensual de la amp litud
térmica diaria. El objetivo de este trabajo es estimar los coeficientes de la ecuación (1) con los dat os obtenidos
para San Nicolás con la finalidad de obtener relaciones preliminares que permitan estimar medias mensuales de
radiación solar diaria en la zona, a partir de los promedios mensuales de amplitud térmica diaria.
METODOLOGÍA
Los datos de irradiancia (I) se tomaron de una estación meteorológica DAVIS Va
ntage Pro2 que registra la
radiación global incidente entre 400 y 1100 nm. Sit uada en la azotea de la Facultad Regional San Nicolás de la
UTN, la estación está dedicada a reportar datos met eorológicos como apoyo de monitoreo ambiental desde
finales del año 2005. La estación estuvo sujeta a calibraciones periódicas; por otra parte, los dato s obtenidos
han sido validados utilizando las facilidades del Centre Energétique et Procédés du Ecole des Mines de Par ís
(www.helioclim.net) (Geiger et al., 2002). Los datos de I se obtuvieron para cada hora. La estación entrega el
máximo (IM) y el promedio (Im) horario en Wm-2. La radiación solar diaria Q se calculó de acuerdo con la siguiente
expresión:
Q = ∫ I dt (2)
la variable t expresa el tiempo de horas de brillo del Sol por d ía (h/d); la integral se estimó utilizando
Im, y se
extiende entre la hora de salida y de puesta del Sol. Para evitar posibles errores en las mediciones debido a la
posición del Sol muy cercana del horizonte y para u na compatibilidad con trabajos futuros, donde pueda aplicarse
la expresión de Angstr öm – Prescott, se ha decidido
utilizar las medidas de irradiancia para las cuales Im > 120 W
m-2 de acuerdo con las bien conocidas prescripciones de la World Meteorological Organization. Como medida de
seguridad, se eliminaron la totalidad de las observaciones para aquellos días en que se detectó algún problema en la
tomas de datos de I. De esta forma, el número total de valores calculados de Q usados para el ajuste fue de 1011.
Los valores Qt se calcularon a partir del instante de salida y puesta solar para cada d ía, considerando la posición
del Sol por sobre 5º de altura a fin de computar horas de irradiancia efectiva, siguiendo los procedimientos
estándar (ver por ejemplo, Meza y Varas, 2000). Se ha tenido en cuenta además la corrección estacional (por
excentricidad de la órbita terrestre) a la “constan te solar ” I0, cuyo valor medio adoptado para los tres años de
estudio fue de 1365,5 W m-2 (Fr öhlich , 2009): este valor se modula con la distancia Tierra-Sol (r) de tal forma
que I0 = 1365,5 W m-2 /(r/UA)2, donde UA es la unidad astronómica; los valores de r se obtuvieron a partir de
integraciones orbitales precisas (Cionco et al., 2012). En particular para el cálculo de la declinació n solar (d) se
ha usado:
-=d
46,23
o
cos
+
d
23,0
π2
64,371
(3)
donde d es un entero que representa el d ía del año (Cionco et al., 2007). Respecto a la amplitud térmica, se
usaron los datos de temperatura para los mismos d ía s considerados en la estimación de Q. Todos los cálculos se
realizaron en ForTran 77 estándar en un entorno Linux, desarrollándose la casi totalidad de los programas
utilizados.
RESULTADOS
La Fig. 1 muestra los valores calculados de Q y los valores te óricos Qt para cada d ía considerado. La evolvente
superior de los puntos graficados, que es una estimación del coeficiente K de d ías claros (sin nubosidad), es
aproximadamente 0,79 x Qt, lo cual está de buen acuerdo con lo reportado por Podestá et al. (2004) para
Pergamino. La totalidad de los valores de Q, oscilan entre 0,3 y 36 MJ m-2 d-1. El cociente de los promedios
mensuales <Q>/<Qt> (36 valores en total, correspondiente a los tres años de estudio) y las ra íces cuadradas de
las amplitudes térmicas promedio se muestran en la Fig. 2. El resultado del ajuste por cuadrados mínim os arroja
a = 0,110; b = 0,181 oC-1, con una valor RMS residual de 0,066 MJ m-2 d-1 representando un RMS porcentual
(RMS%) de 9,690 %.
El RMS residual se define como:
(సభ ି )మ
RMS = ට∑
మ
(4)
donde Koi son los cocientes <Q>/<Qt> “observados”, esto es, utilizados para el ajuste
y calculados según lo
indicado en la sección precedente; Kci son los valores <Q>/<Qt> obtenidos a partir del ajuste; n es el número
total de datos utilizados en el ajuste. El RMS% se obtiene dividiendo el RMS por el promedio de los Koi
utilizados en el ajuste y multiplicando por 100.
Figura 1: Valores observados de radiación solar dia ria utilizados para el ajuste (puntos) y valores teóricos por fuera de la
atmósfera (l ínea punteada).
Figura 2: Recta de regresi ón ajustada al cociente d e los promedios mensuales de Q y Qt en función de l a ra íz cuadrada de la
amplitud térmica mensual promedio. El coeficiente de correlación R es 0,848.
Figura 3: Residuos post-ajuste en función del mes c onsiderado. Se observan algunos efectos sistemáticos.
De esta forma, la estimación de la media mensual de Q para un mes determinado, se obtiene hallando los
promedios mensuales de amplitud térmica y de Qt, introduciendo estos valores en la ecuación (1) co n los
coeficientes ajustados:
<Q> = <Qt> (0,110 + 0,181 oC-1 <DT>1/2 ). (5)
La Fig. 3, muestra los residuos post-ajuste. Los valores absolutos de los residuos son siempre menores que 0,15
WJ m-2 d-1, manteniéndose bien por dentro del límite ± 3 RMS. Se observan algunos efectos sistemáticos:
mientras que los primeros 12 valores (primer año) se distribuyen con bastante normalidad, los residuos
correspondientes a los otros años (del 12 al 24 y del 25 al 36) muestran sistemáticamente mayor número de
residuos positivos (segundo año) y negativos (tercer año). Sin embargo, la distribución de los residuo s dentro del
límite señalado, indica que los efectos sistemáticos no son importantes (no existen valores “outliers”
significativos; Arley, 1950; Bevington, 1968).
COMENTARIOS FINALES Y CONCLUSIONES
En este trabajo se han determinado los coeficientes de un modelo lineal que relaciona valores mensuales
promedio de Q y la ra íz cuadrada de medias mensuales correspondi entes de amplitud térmica diaria, en base a
una serie de tres años de datos de irradiancia solar, obtenida en la Facultad Regional San Nicolás de UTN. Los
resultados indican que la utilización de la ecuació
n (1) para establecer dicha relación es muy razonab le y su
estudio en la zona debe profundizarse con más datos. Al respecto, es importante mencionar que se han ensayado
otras relaciones posibles, distintas a la ecuación (1); por ejemplo, la relación de Bristow-Campbell ( Bristow y
Campbell, 1984), o el ajuste de un coeficiente distinto a ½ en la potencia de <DT>, pero los resultados
decididamente fueron de menor calidad estad ística.
Por último, los resultados obtenidos están de buen acuerdo con lo reportado por Raichijk et al. (2005, tabla 1),
obteniéndose valores concordantes con los de Paraná, Entre Ríos, a unos 150 km al norte de San Nicolás. El
RMS % obtenido indica que el error esperado en la predicción de los valores de Q también está dentro de lo
estimado por esos autores. Los efectos sistemáticos presentes en los residuos deben estudiarse en función de un
mayor número de observaciones.
AGRADECIMIENTOS
Este trabajo se ha realizado en el marco del proyecto PID-UTN 1351 “Forzantes externos al planeta y
variabilidad climática ”.
REFERENCIAS
Ångstr öm, A. (1924). Solar and terrestrial radiatio n. Q. J. R. Meteorol. Soc. 50, 121 –125.
Arley N. y Buch P. (1950). Introduction to the theory of probability and statistics. New York: Wiley
Bandyopadhyay A., Bhadra A., Raghuwanshi N.S. y Singh R. (2008). Estimation of monthly solar radiation
from measured air temperature extremes. Agricultural and forest meteorology 148, 1707 –1718.
Bevington P.R. (1968). Data reduction and error analysis for the physical sciences. New York: McGraw-Hill.
Bristow, K.L. y Campbell, G.S. (1984). On the relationship between incoming solar radiation and daily
maximum and minimum temperature. Agric. For. Meteorol. 31, 159 –166.
Cionco R., Caligaris M. y Quaranta N. (2007). Herramientas de cálculo de la orientación solar. Modeliz ación
Aplicada a la Ingenier ía, Volumen II. Editores: Wal ter E. Legnani, Pablo Jacovkis, Ricardo L. Armentano.
Ed. UTN, Fac. Regional Bs. As.
Cionco R., Quaranta N., Caligaris M. y Compagnucci, R. (2012). Modelado de alta precisión en el cálcul o de la
orientación solar. Revista Tecnología y Ciencia 20,
51-53. ISSN 1666-6917 Ed. UTN.
Fr öhlich C. ( 2009). Evidence of a long-term trend in, total solar irradiance. Atronomy & Astrophysics, 501, L27-
L30.
Geiger M., Diabate´ L., Me´nard L. y Wald, L. (2002 ). A web service for controlling the quality of
measurements of global solar irradiation. Solar Energy 73, 6, 475 –480.
Meinke H., Carberry P.S., McCaskill M.R., Hills M.A., McLeod I. (1995). Evaluation of radiation and
temperature data generators in the Australian tropics and sub-tropics using crop simulation models. Agric.
For. Meteorol. 72, 295–316.
Meza F. y Varas E. (2000). Estimation of mean monthly solar global radiation as a function of temperature.
Agricultural and Forest Meteorology 100, 231 –241.
Podestá G., Núñez L., Villanueva C. y Skansi M. (2004). Estimating daily solar radiation in the Argentine
Pampas. Agricultural and Forest Meteorology, 123, 41–53.
Prescott, J.A. (1940). Evaporation form a water surface in relation to solar radiation. Trans. R. Soc. Sci. Aust. 64,
114 –125.
Raichijk C., Grossi Gallegos H. y Righini R. (2005). Evaluación de método alternativo para la estimaci ón de
valores medios mensuales de irradiación global en a rgentina. Avances en Energías Renovables y Medio
Ambiente, 9, 11.05-11.08.
Raichijk C. y Grossi Gallegos H. (2007). Evaluación de una aplicación alternativa de la fórmula de Sue
la república argentina. Avances en Energías Renovab les y Medio Ambiente, 11, 11.25-11.30.
hrcke en
Righini R. y Grossi Gallegos H. (2003). Aproximació n a un trazado de nuevas cartas de irradiación sola r para
argentina. Avances en Energías Renovables y Medio A mbiente, 7, 2, 11.07-11.11.
Rusticucci M. y Vargas W. (2002). Cold and warm events over Argentina and their relationship with the ENSO
phases risk evaluation analysis. Int. J. Climatol. 22, 467 –483.
Seinfeld J. H. (2001). Atmospheric Chemistry and Physics of Air Pollution. John Wiley & Sons.
|
1906.10330 | 1 | 1906 | 2019-06-25T05:55:19 | Thermal Tides in Rotating Hot Jupiters | [
"astro-ph.EP"
] | We calculate tidal torque due to semi-diurnal thermal tides in rotating hot Jupiters, taking account of the effects of radiative cooling in the envelope and of the planets rotation on the tidal responses. We use a simple Jovian model composed of a nearly isentropic convective core and a thin radiative envelope. To represent the tidal responses of rotating planets, we employ series expansions in terms of spherical harmonic functions $Y_l^m$ with different $l$s for a given $m$. For low forcing frequency, there occurs frequency resonance between the forcing and the $g$- and $r$-modes in the envelope and inertial modes in the core. We find that the resonance enhances the tidal torque, and that the resonance with the $g$- and $r$-modes produces broad peaks and that with the inertial modes very sharp peaks, depending on the magnitude of the non-adiabatic effects associated with the oscillation modes. We also find that the behavior of the tidal torque as a function of the forcing frequency (or period) is different between prograde and retrograde forcing, particularly for long forcing periods because the $r$-modes, which have long periods, exist only on the retrograde side. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000 -- 000 (2019)
Printed 26 June 2019
(MN LATEX style file v2.2)
Thermal Tides in Rotating Hot Jupiters
Umin Lee⋆ and Daiki Murakami†
Astronomical Institute, Tohoku University, Sendai, Miyagi 980-8578, Japan
Accepted XXX. Received YYY; in original form ZZZ
ABSTRACT
We calculate tidal torque due to semi-diurnal thermal tides in rotating hot Jupiters,
taking account of the effects of radiative cooling in the envelope and of the plan-
ets rotation on the tidal responses. We use a simple Jovian model composed of a
nearly isentropic convective core and a thin radiative envelope. To represent the tidal
responses of rotating planets, we employ series expansions in terms of spherical har-
monic functions Y m
l with different ls for a given m. For low forcing frequency, there
occurs frequency resonance between the forcing and the g- and r-modes in the enve-
lope and inertial modes in the core. We find that the resonance enhances the tidal
torque, and that the resonance with the g- and r-modes produces broad peaks and
that with the inertial modes very sharp peaks, depending on the magnitude of the
non-adiabatic effects associated with the oscillation modes. We also find that the be-
havior of the tidal torque as a function of the forcing frequency (or period) is different
between prograde and retrograde forcing, particularly for long forcing periods because
the r-modes, which have long periods, exist only on the retrograde side.
Key words: hydrodynamics - waves - stars: rotation - stars: oscillations - planet
-star interactions
1 INTRODUCTION
It is well known that equilibrium and dynamical gravitational tides play various important roles in the dynamics of planetary
systems and in binary systems of stars (e.g., Goldreich & Soter 1966; Zahn 1977; Press & Teukolsky 1977; Savonije &
Papaloizou 1984, 1997; Lai 1997; Witte & Savonije 2002; Ivanov & Papaloizou 2007; Ogilvie & Lin 2004; Ogilvie 2014; Fuller
& Lai 2013). Because of energy dissipations accompanied by the tidal responses raised in stars and planets, the gravitational
tides can be a cause of synchronization between the rotation and the orbital motion, and of circularization of the orbits.
Hot Jupiters are giant gas planets orbiting the central star at distances as close as less than about 0.05A.U.. Observa-
tionally it is suggested that hot Jupiters have larger radii compared to Jovian planets of the similar mass and age orbiting
far from the central stars (e.g., Jermyn, Tout & Ogilvie 2017). Baraffe et al (2003), for example, have numerically shown
that the planets would inflate to the extent observationally determined if there exists an effective heating source deep in the
atmosphere. As a possible heating mechanism, tidal heating in the planets was suggested by Bodenheimer, Lin, Mardling
(2001). In the case of hot Jupiters closely orbiting the central star, however, the gravitational tides are strong enough to
make the spin and orbital motion synchronized in a timescale much shorter than the ages of the planets, which suggests that
another mechanism is needed that keeps the spin and orbital motion asynchronous. For such a mechanism for hot Jupiters,
Arras & Socrates (2010) suggested thermal tides, which would be operative because of strong irradiation by the central star.
The periodic alternations of day and night sides on the planets may bring about semi-diurnal density perturbations in the
planets, which could cancel the effects of the density perturbations produced by the gravitational tides. If this is the case,
the spin and orbital motion would be kept asynchronous so that the tidal heating can be a viable mechanism to inflate the
planets (e.g., Arras & Socrates 2010).
Following the suggestion by Arras & Socrates (2010), Auclair-Desrotour & Leconte (2018) computed the tidal torque
caused by semi-diurnal thermal tides, taking into considerations the effects of energy dissipations caused by radiative cooling
⋆ e-mail:[email protected]
† e-mail:[email protected]
c(cid:13) 2019 RAS
2
U. Lee, D. Murakami
in the envelope and those of the planets rotation on the tidal responses. For hot Jupiters, they used very simple models,
composed of a convective core and a thin radiative envelope. The convective core has the structure corresponding to that of
a polytrope of the index n = 1 and the envelope is nearly isothermal. To represent the tidal responses in rotating planets,
Auclair-Desrotour & Leconte (2018) employed series expansions in terms of the Hough functions, defined in the traditional
approximation for the perturbations in rotating bodies (e.g., Lee & Saio 1997). They found that the tidal torque due to
thermal tides is affected by frequency resonance between the tidal forcing and the g-modes in the envelope, changing its sign
as a function of the forcing period. They also discussed the possibility to produce differential rotation at the sites where the
tide exerts strong local torques.
We revisit the problems of the semi-diurnal thermal tides in rotating hot Jupiters, following Auclair-Desrotour & Leconte
(2018). In this paper, however, we do not employ the traditional approximation to represent tidal responses in the rotating
planets. Instead, we simply use series expansions in terms of spherical harmonic functions Y m
l (θ, φ) with different ls for a
given m (see, e.g., Lee & Saio 1986, 1987). We also assume the convective core of the rotating planets is nearly isentropic (e.g.,
Stevenson & Salpeter 1977a,b; Stevenson 1979) so that the core can support propagation of inertial modes. In this paper, we
discuss for rotating Jovian planets tidally excited low frequency modes such as g-modes, inertial modes and r-modes. The
g-modes are internal gravity waves propagating in stably stratified regions expected for the radiative envelope of the planet.
The latter two are rotationally induced modes for which the Coriolis force is the restoring force. There exist prograde and
retrograde inertial modes but r-modes exist only on the retrograde side (Papaloizou & Pringle 1978; see also Greenspan 1969;
Unno et al 1989). r-modes result from the interaction between the radial component of the vorticity and the Coriolis force
and form a subclass of inertial modes. Note that although inertial modes propagate in nearly isentropic regions, the r-modes
discussed in this paper are those propagating in the radiative envelope. See also the 5th paragraph in §3. The method of
solution used in this paper is described in §2 and the numerical results are given in §3. We conclude in §4.
2 BASIC EQUATIONS
2.1 Equilibrium Model
Following Arras & Socrates (2010) and Auclair-Desrotour & Leconte (2018), to compute tidal responses of strongly irradiated
Jovian planets, we use a simple model, composed of an irradiated thin isothermal atmosphere and a nearly isentropic convective
core. Such Jovian models are obtained by integrating the equations for hydrostatic equilibrium (e.g., Clayton 1968)
dp
dr
= −ρ
GMr
r2
,
dMr
dr
= 4πr2ρ,
with the analytic equation of state given by
ρ(p) = e−p/pb p
a2 +(cid:16)1 − e−p/pb(cid:17)r p
,
Kc
where p, ρ, Mr, and G are respectively the gas pressure, the mass density, the mass within the sphere of radius r and the
gravitational constant, and pb is the pressure at the base of the stably stratified layer (radiative atmosphere) and
(1)
(2)
(3)
(4)
(5)
(6)
Kc = GR2
J ,
a2 =ppbKc,
where a is the isothermal sound velocity and RJ is the radius of Jupiter. In the convective core where p ≫ pb, we obtain
p ≈ KcρΓ with Γ = 2,
and hence the square of the Brunt-Vaisala frequency
N 2 = −g(cid:18) d ln ρ
dr
−
1
Γ1
d ln p
dr (cid:19) ≡ −gA ≈ 0
for Γ1 ≡(cid:18) ∂ ln p
∂ ln ρ(cid:19)ad
= 2,
where g = −GMr/r2, and A is known as the Schwarzschild discriminant. On the other hand, in the radiative layers where
p ≪ pb, we have
p ≈ a2ρ,
and
N 2 ≈ (Γ1 − 1)
g2
c2 ,
c2 = Γ1
p
ρ
,
(7)
(8)
and c is the adiabatic sound speed. Since p/ρ ≈ a2 ∝ T for an ideal gas, the temperature T is constant for a constant a,
suggesting an isothermal atmosphere.
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
2
L
l
0
1
0
1
g
o
l
&
N
g
o
l
2
Thermal Tides in Hot Jupiters
3
4
2
4
2
0
-2
-4
12
10
8
log
6
p
10
Figure 1. Square of the Brunt-Vaisala frequency N (solid line) and the Lamb frequency Ll for l = 2 (dotted line) plotted as a function
of log10 p for the Jovian model, where N 2 and L2
l are normalized by GM/R3 with M and R being the mass and radius of the model.
In this paper, we use pb = 100bar = 108dyn/cm2 and set the outer boundary Re at p = 0.01dyn/cm2 and define the
planet's radius R = Re/1.01. We use M = 0.7MJ with MJ being the mass of Jupiter and the radius R = 9.31 × 109cm.
l ≡ l(l + 1)c2/r2, normalized by GM/R3,
Figure 1 is the propagation diagram of the Jovian model we use, where N 2 and L2
are plotted for l = 2 versus log10 p. Note that Ll is called Lamb frequency and indicates the local lower limit of frequency for
the propagation of sound waves.We have N 2 ≈ 0 for Γ1 = 2 in the convective core, and N 2 > 0 in the radiative envelope.
Since g-modes of the frequency ω propagate in the regions where ω < N and ω < Ll, Figure 1 indicates that the g-modes
that propagate in the radiative envelope of the model have frequencies ω . 0.1pGM/R3.
2.2 Perturbed Basic Equations
the angular speed Ω and is orbiting the host star in a circular orbit with the frequency Ωorb = p(M∗ + M )/a3
We treat tidal responses as small amplitude perturbations of the planet. We assume that the planet is uniformly rotating at
∗, where a∗
denotes the semi-major axis, and M∗ and M are the mass of the host star and the planet, respectively. Assuming that the
perturbing tidal potential ΦT due to the host star depends on the time as ΦT ∝ eiωt with ω being the forcing frequency, the
basic equations for the small amplitude tidal responses of the planet may be governed by (see, e.g., Unno et al 1989)
−ρω2ξ + 2iρωΩ × ξ −
ρ′
ρ
∇p + ∇p′ = −ρ∇ΦT ,
ρ′ + ∇ · (ρξ) = 0,
iωρT δs = (ρǫ)′ − ∇ · F ′,
δs
cp
=
1
αT (cid:18) 1
Γ1
δp
p
−
δρ
ρ (cid:19) ,
(9)
(10)
(11)
(12)
where s, T , cp, ǫ, and F are respectively the specific entropy, temperature, specific heat at constant pressure, energy generation
rate per gram, and energy flux vector,and αT = − (∂ ln ρ/∂ ln T )P is the volume expansion coefficient, ∇ad = (∂ ln T /∂ ln p)s
is the adiabatic temperature gradient, and ξ is the displacement vector, and (′) and δ indicate respectively the Eulerian and
Lagrangian perturbations. Note that vectorial quantities are indicated by using italic bold faces. Equations (9), (10), (11),
and (12) are perturbed versions of the equation of motion, continuity equation, entropy equation and the equation of state,
respectively. Note that we have applied the Cowling approximation, neglecting the Euler perturbation of the gravitational
potential due to self-gravity, and that we have ignored the rotational deformation of the planet so that the equilibrium
structure is spherical symmetric.
For the entropy equation (11), we employ the approximation used by Auclair-Desrotour & Leconte (2018). Assuming
the Newtonian cooling (e.g., Mihalas & Mihalas 1999), the second term on the right-hand-side of equation (11) may be
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
4
U. Lee, D. Murakami
approximately given by
∇ · F ′ ∼ ωDρcpT
T ′
T
= ωDρcpT (cid:18) δs
cp
+ ∇ad
δp
p
+ ∇V
ξr
r (cid:19) ,
where V = −d ln p/d ln r, ∇ = d ln T /d ln p,
ωD =
4π
τ∗ "(cid:18) p
p∗(cid:19)1/2
+(cid:18) p
p∗(cid:19)2#−1
,
(13)
(14)
and p∗ is the pressure at the base of the heated layer, τ∗ is the timescale parameter to specify the efficiency of radiative cooling
in the envelope, and we use p∗ = 106dyn/cm2 (see Auclair-Desrotour & Leconte 2018; Iro et al 2005). Note that to derive
equation (13) we have used the relations given by δs/cp = δT /T − ∇adδp/p and δT /T = T ′/T + ξrd ln T /dr. Substituting
equation (13) into equation (11), we obtain the entropy perturbation δs given by
δs
cp
=
1
iω + ωD
(ρǫ)′
ρT cp
−
ωD
iω + ωD (cid:18)∇ad
δp
p
+ ∇V
ξr
r (cid:19) .
(15)
In this paper, we assume uniform rotation for simplicity. The convective core is likely to rotate uniformly if turbulent
mixing is efficient enough in the core. If this is the case, uniform rotation can be a good approximation since most of the
mass of hot Jupiters is occupied by the convective core. If there exists a strong differential rotation between the core and the
envelope, the shear would excite turbulence, which could weaken the differential rotation if the buoyant effects are weak (e.g.,
Turner 1979). If there consistently exists a differential rotation between the core and the envelope, the frequency spectra of
tidally excited g-modes and inertial modes would be different from those found assuming uniform rotation since the frequency
of inertial modes is simply proportional to the rotation rate of the core but the frequency of g-modes is not.
2.3 Forced Oscillation Equations
In the presence of gravitational and/or thermal tidal forcing, the equations that govern the tidal responses become a set of
inhomogeneous linear differential equations. Assuming the rotation axis of the planet is perpendicular to the orbital plane,
the tidal potential in the planet due to the host star may be given by
ΦT = −
GM∗
r − a∗(t)
= −GM∗Xlm
Wlm
rl
a∗(t)l+1 e−imΦY m
l (θ, φ),
(16)
where the origin of spherical polar coordinates (r, θ, φ) is at the centre of the planet, a∗(t) is the position vector to the host
star, a∗(t) = a∗(t), Φ is the true anomaly, and
Wlm = (−1)(l+m)/2(cid:20) 4π
2l + 1
(l − m)!(l + m)!(cid:21)1/2
2 (cid:19)!(cid:21)
2 (cid:19)!(cid:18) l + m
(cid:20)2l(cid:18) l − m
= Wl,−m,
(17)
which has a non-zero value for even values of l − m. Note that the tidal potential ΦT does not contain the dipole terms
associated with l = 1. Assuming that the eccentricity e of the orbit is small so that Φ ≈ Ωorbt, and taking only the dominant
tidal component with l = −m = 2, the tidal potential ΦT in an inertial frame may be given by
ΦT = −W2,−2
GM∗
a3
∗
r2Y −2
2
(θ, φ)e2iΩorbt.
(18)
When the planet is uniformly rotating at Ω, the tidal potential in the co-rotating frame may be given by replacing φ by φ + Ωt
and hence the forcing frequency 2Ωorb by ω = 2Ωorb + mΩ = 2(Ωorb − Ω) for m = −2.
Thermal tides are caused by insolation by the host star, which produces the day and night sides on the planet and is
given by (e.g., Auclair-Desrotour & Leconte 2018)
ǫ′ =( J∗(r, θ, φ, t) = κ∗F∗e−p(r)/p∗ cos φ∗
J∗(r, θ, φ, t) = 0
for
0 6 φ∗ 6 π/2
for π/2 6 φ∗ 6 π,
where φ∗ is the zenith angle of the host star as observed from the planet and is given by cos φ∗ = sin θ cos(φ − Φ), and
F∗ = σSBT 4
r∗ (cid:19)2
∗ (cid:18) R∗
,
κ∗ =
g0
p∗
,
g0 =
GM
R2 ,
(19)
(20)
and T∗ and R∗ are respectively the surface temperature and radius of the host star, κ∗ is the opacity at the base of the heated
layer and r∗ is the distance between the planet and the host star, set equal to the semi-major axis a∗ of the orbit. Assuming
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
Thermal Tides in Hot Jupiters
5
Φ = Ωorbt, we obtain
Z π
0
dθ sin θZ Φ+π/2
Φ−π/2
dφ(Y −2
2
)∗ cos φ∗ =
1
16r 15π
2
e2iΩorbt.
In general, in the co-rotating frame of the planet, we may expand J∗ in terms of spherical harmonic function as
J (lmn)
∗
(r)Y m
l (θ, φ)eiωmnt,
J∗ = Xl,m,n
(21)
(22)
where ωmn = nΩorb + mΩ and nΩorb represents the forcing frequency in an inertial frame. If we take only the component
with n = −m = 2, for which ω−2,2 = 2(Ωorb − Ω), assuming (ρǫ)′ = ρǫ′ since the unperturbed state has no insolation so that
ǫ = 0, we obtain
δs
cp
=
1
iω + ωD
ǫ′
T cp
−
ωD
iω + ωD (cid:18)∇ad
δp
p
+ ∇V
ξr
r (cid:19) ,
where ω = ω−2,2 and
ǫ′ =
1
16r 15π
2
κ∗F∗e−p(r)/p∗Y −2
2
(θ, φ)eiωt.
(23)
(24)
To represent the perturbations of tidally perturbed and rotating planets, we employ series expansion in terms of spherical
harmonic functions Y m
l (θ, φ) for a given m with different ls (e.g., Lee & Saio 1986). The pressure perturbation is given by
p′(r, θ, φ, t) =Xl
l(r)Y m
p′
l (θ, φ)eiωt,
and the displacement vector ξ by
Sl(r)Y m
l (θ, φ)eiωt,
ξr(r, θ, φ, t) = rXl
ξθ(r, θ, φ, t) = rXl,l′ (cid:20)Hl(r)
ξφ(r, θ, φ, t) = rXl,l′ (cid:20)Hl(r)
∂
∂θ
Y m
l (θ, φ) + Tl′
1
sin θ
∂
∂φ
1
sin θ
∂
∂φ
Y m
l (θ, φ) − Tl′
∂
∂θ
Y m
l′ (θ, φ)(cid:21) eiωt,
l′ (θ, φ)(cid:21) eiωt,
Y m
(25)
(26)
(27)
(28)
j = lj + 1 for even modes, and lj = m + 2j − 1 and l′
where lj = m + 2(j − 1) and l′
j = lj − 1 for odd modes and
· · · , jmax. In this paper, since we assume for simplicity that the spin axis of the planet is perpendicular to
j = 1, 2,
the orbital plane of the planet and that the planet is on a circular orbit around the host star, we retain only the forcing
terms proportional to Y −2
2 eiωt with ω = 2(Ωorb − Ω), for which the tidal responses may be represented by the sum of terms
proportional to Y −2
· · · , jmax. Note that jmax determines the length of the series expansions
for the responses, and we use jmax = 12 in this paper.
eiωt with l = 2j and j = 1, 2,
l
Substituting the expansions into the perturbed basic equations, we obtain a set of linear ordinary differential equations
with inhomogeneous terms. Defining the dependent variables as
y1 = (Slj ), y2 = p′
lj
ρgr! , y6 =(cid:18) δslj
cp (cid:19) , h = (Hlj ),
t = (Tl′
j
), Y 2 = y2 +
ψ
gr
,
the perturbed basic equations reduce to
r
dy1
dr
=(cid:18) V
Γ1
− 3(cid:19) y1 −
V
Γ1
Y 2 + Λ0h + αT y6 +
V
Γ1
ψ
gr
,
r
dY 2
dr
= (c1 ¯ω2 + rA)y1 + (1 − U − rA)Y 2 − 2c1 ¯ω ¯Ω (mh + C0it) + αT y6 + rA
ψ
gr
,
−M0h + L1it = −νKy1,
L0h − M1it = mνΛ−1
0 y1 +
1
c1 ¯ω2
Y 2,
y6 =
1
iω + ωD
j ∗ −
ωD
iω + ωD
V (cid:20)∇adY 2 − ∇ad
ψ
gr
+ (∇ − ∇ad) y1(cid:21) ,
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
(29)
(30)
(31)
(32)
(33)
(34)
6
U. Lee, D. Murakami
where U = d ln Mr/d ln r, c1 = (r/Rp)3/(Mr/M ), and
ν =
2Ω
ω
,
¯ω =
ω
σ0
,
¯Ω =
Ω
σ0
,
σ0 =r GM
R3 ,
and non-zero elements of the matrices Λ0, C0, L0, L1, K, M0, M1 for even modes are defined by
(Λ0)j,j = lj (lj + 1),
(C0)j,j = −(lj + 2)J m
lj +1,
(C0)j+1,j = (lj + 1)J m
lj +2
(L0)j,j = 1 −
mν
lj(lj + 1)
,
(L1)j,j = 1 −
mν
j + 1)
l′
j (l′
,
(K)j,j =
J m
lj +1
lj + 1
,
(K)j,j+1 = −
J m
lj +2
lj + 2
,
(M0)j,j = ν
lj
lj + 1
J m
lj +1,
(M0)j,j+1 = ν
lj + 3
lj + 2
J m
lj +2,
(M1)j,j = ν
lj + 2
lj + 1
J m
lj +1,
(M1)j+1,j = ν
lj + 1
lj + 2
J m
lj +2.
(35)
(36)
(37)
(38)
See the Appendix A for the derivation of equations (30) to (34). The vectors j ∗ and ψ are inhomogeneous forcing terms and
have only the first component given respectively by
(j ∗)1 = p15π/2
16
κ∗F∗
T cp
e−p(r)/p∗,
and
(ψ)1
gr
=
ΦT
gr
= −r 3π
10
M∗
M∗ + M
c1 ¯Ω2
orb
with ¯Ωorb = Ωorb/σ0. To estimate the temperature T and the specific heat cp, we assume those for an ideal gas, that is,
T =
µ
R
p
ρ
,
cp =
5
2
R
µ
,
and R is the gas constant, and µ is the mean molecular weight, for which we use µ = 1.3.
Using the auxiliary equations (32) and (33), we obtain
Λ0h =
W
c1 ¯ω2
Y 2 + νWOy1,
2c1 ¯ω ¯Ω(mh + C0it) = νOT WY 2 + 4c1 ¯Ω2Gy1,
where
W = Λ0(L0 − M1L−1
1
M0)−1, O = mΛ−1
0 − M1L−1
1
K, G = OT WO − C0L−1
1
K.
Substituting equations (42), (43), and (34), we obtain the set of linear differential equations for forced oscillations:
r
dy1
dr
r
dY 2
dr
Γ1
− 3 + β1(cid:19) 1 + νWO(cid:21) y1 +(cid:20) W
=(cid:20)(cid:18) V
=(cid:2)(cid:0)c1 ¯ω2 + rA + β1(cid:1) 1 − 4c1 ¯Ω2G(cid:3) y1 +h(1 − U − rA − β2) 1 − νOT Wi Y 2 +
+ β2(cid:19) 1(cid:21) Y 2 +
c1 ¯ω2 −(cid:18) V
iω + ωD
αT
Γ1
where 1 is the unit matrix, and
j ∗ +(cid:18) V
Γ1
+ β2(cid:19) ψ
gr
,
αT
iω + ωD
j ∗ + (rA + β2)
ψ
gr
,
β1 =
αT (∇ad − ∇)V
iω/ωD + 1
,
β2 =
αT ∇adV
iω/ωD + 1
.
The boundary condition at the centre is the regularity condition of the functions y1 and Y 2 (see the Appendix B). The outer
boundary condition at the surface of the planet is given by δp = 0 (see, e.g., Unno et al 1989).
3 NUMERICAL RESULTS
The tidal torque on the planet may be given by (e.g., Auclair-Desrotour & Leconte 2018)
(39)
(40)
(41)
(42)
(43)
(44)
(45)
(46)
(47)
(48)
N = −Z ∂ΦT
∂φ
ρ′dV = −
1
2 Z Re(cid:18) ∂ΦT
∂φ
ρ′∗(cid:19) dV = −Z Im(cid:0)ΦT ρ′∗(cid:1) dV,
where ρ′ is the tidal response caused by the tides associated with ΦT and/or ǫ′. If the density perturbation in the rotating
planet is represented by the series expansion similar to (25) and the tidal potential ΦT (r, θ, φ) is simply proportional to
Y −2
(θ, φ) as given by equation (18), the tidal torque N is computed by
2
N =r 3π
10
GM∗
∗ Z R
0
a3
drr4Im[ρ′∗
2 (r)],
(49)
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
N
g
o
l
0
1
32
31
30
29
28
27
26
0
Thermal Tides in Hot Jupiters
7
32
31
30
29
28
27
26
N
g
o
l
0
1
0.05
0.1
0.15
0.2
0
0.05
0.1
0.15
0.2
ω/σ
0
ω/σ
0
Figure 2. Tidal torque, given in erg, due to thermal tides (left panel) and gravitational tides (right panel) as a function of the forcing
frequency ¯ω = ω/σ0 for ¯Ω = 0, where the red (black) lines are for positive (negative) N , and the dash-dotted lines, solid lines, and dotted
lines are for τ∗ = 0.1, 1, and 10 days, respectively.
Table 1. Complex eigenfrequency ¯ω = ¯ωR + i¯ωI of l = 2 gn-modes in the envelope for ¯Ω = 0.
τ∗ (day)
0.1
1
10
n
1
2
3
4
¯ωR
¯ωI
¯ωR
¯ωI
¯ωR
¯ωI
9.29 × 10−2
4.15 × 10−2
2.40 × 10−2
1.64 × 10−2
1.36 × 10−2
1.24 × 10−2
8.50 × 10−3
6.33 × 10−3
1.09 × 10−1
5.86 × 10−2
3.62 × 10−2
2.54 × 10−2
1.04 × 10−2
1.24 × 10−2
9.63 × 10−3
7.52 × 10−3
1.24 × 10−1
7.98 × 10−2
5.77 × 10−2
4.25 × 10−2
1.06 × 10−2
1.97 × 10−2
1.96 × 10−2
1.38 × 10−2
where the tidal response ρ′
2(r) is obtained by solving the inhomogeneous linear differential equations (45) and (46) for a given
forcing frequency ω. For numerical computations, we assume M∗ = M⊙, R∗ = R⊙, and T∗ = 5.8 × 103K for the host star,
and the distance between the planet and the star is assumed to be r∗ = a∗ = 0.05A.U..
We first discuss the case of ¯Ω = 0. We calculate non-adiabatic free g-modes propagating in the radiative envelope of the
planet for different values of τ∗, which determines the thermal time scales in the envelope. Free g modes may be computed by
setting ψ = 0 and j ∗ = 0 in equations (45) and (46) and introducing a normalization condition, for example, given by Sl1 = 1
at the surface. Note that non-adiabatic g-modes have complex eigenfrequency ¯ω. The result of non-adiabatic calculation of
free g-modes is summarized in Table 1, in which complex eigenfrequency ¯ω is given for several low radial order g modes for
three values of the parameter τ∗. The shorter τ∗ is, the larger the non-adiabatic effects are. The table shows that the g-modes
are all pulsationally stable and have large damping rate η ≡ ωI/ωR & 0.1, where ωR and ωI are the real and imaginary parts
of the complex frequency ω (see the caption to Table 1). It also shows that the frequency ¯ωR,n of the gn-mode increases but
the frequency difference ¯ωR,n − ¯ωR,n+1 decreases as the radial order n and τ∗ increase.
Figure 2 plots the absolute value of the tidal torque N for ¯Ω = 0 as a function of the forcing frequency ¯ω for three
different values of τ∗, where the left panel is for the case of ψ = 0 and j ∗ 6= 0, and the right panel is for the case of ψ 6= 0 and
j ∗ = 0 in equations (45) and (46). For both cases, there appears, as a function of ¯ω, broad peaks of N , which are produced
by frequency resonance between the forcing frequency and the natural frequency of the g-modes. The width of the peaks
may be determined by the magnitude of τ∗, that is, the larger τ∗ is, the narrower the peaks are, which is partly because the
frequency difference ¯ωR,n − ¯ωR,n+1 decreases with increasing τ∗ for the g-modes with large damping rates η.
The sign of the tidal torque at the resonance peaks alternately changes as the g-mode in resonance with the forcing is
changed with decreasing ¯ωR, except for the case of the gravitational tide for τ∗ = 10day. The response to the tidal potential
ΦT is quite similar to that to the thermal tides except for very low tidal frequency region. The tidal torque N due to the
thermal tides decreases as ¯ω → 0, but the torque due to the gravitational tides tend to a constant value, corresponding to that
of the gravitational equilibrium tide, the magnitudes of which is proportional to τ −1
∗ , that is, the possible amount of energy
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
8
U. Lee, D. Murakami
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
-1
-0.5
0
0.5
1
-0.2
-0.1
0
0.1
0.2
ω/σ
0
ω/σ
0
Figure 3. Tidal torque, given in erg, due to thermal tides (j ∗ 6= 0 and ψ = 0) as a function of the forcing frequency ¯ω for ¯Ω = 0.05,
where positive (negative) ¯ω indicates prograde (retrograde) forcing, and the red (black) lines are for positive (negative) N .
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
-1
-0.5
0
0.5
1
-0.2
-0.1
0
0.1
0.2
ω/σ
0
ω/σ
0
Figure 4. Same as Figure 3 but for ¯Ω = 0.1.
dissipation in the envelope. Since we consider no effects of turbulent fluid motion in the convective core on tidal responses
due to the gravitational tidal potential ΦT , we have to be cautious about estimating the tidal effects on the planets.
Let us briefly discuss the modal properties of low frequency, even parity, free oscillation modes of the rotating planets,
such as g-modes, r-modes and inertial modes. Oscillation modes of rotating planets are separated into prograde modes and
retrograde modes, observed in the co-rotating frame of the planet. In our convention, for negative m, prograde (retrograde)
modes correspond to positive (negative) spin parameter ν = 2Ω/ω where ω is the oscillation frequency observed in the co-
rotating frame of the planet. The modal properties of low frequency g-modes, frequency and stability, are affected by rotation,
particularly when ν & 1. Rotation also produces new kinds of oscillation modes, called inertial modes and r-modes. Note
that inertial modes propagate in nearly isentropic regions and that r-modes, which form a subclass of inertial modes, appear
only as retrograde modes. The restoring force for inertial modes is the Coriolis force, and their frequency ω is proportional to
the rotation frequency Ω and the ratio ω/Ω is limited to ω/Ω 6 2. In other wards, inertial modes appear when ν > 1. The
radiative envelope is the propagation regions of r-modes of even parity, for which both Coriolis force and buoyant force plays
essential roles. It is well known that the asymptotic frequency ω of r-modes in the limit of Ω → 0 is given by 2mΩ/[l′(l′ + 1)].
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
Table 2. Complex eigenfrequency ¯ω = ¯ωR + i¯ωI of gn-modes, rn-modes and inertial modes iL with L ≡ l0 − m for m = −2 and ¯Ω = 0.1,
where we use τ∗ =1 day.
prograde
retrograde
Thermal Tides in Hot Jupiters
9
modes
¯ωR
9.36 × 10−2
4.85 × 10−2
2.95 × 10−2
· · ·
· · ·
· · ·
¯ωI
9.47 × 10−3
1.07 × 10−2
8.05 × 10−3
· · ·
· · ·
· · ·
5.55 × 10−2
1.27 × 10−1
2.74 × 10−2
1.16 × 10−5
9.89 × 10−6
3.35 × 10−6
g1
g2
g3
r1
r2
r3
i2
i4
i4
¯ωR
−1.68 × 10−1
−1.23 × 10−1
−9.91 × 10−2
−2.07 × 10−2
−1.19 × 10−2
−7.47 × 10−3
−1.10 × 10−1
−1.52 × 10−1
−8.60 × 10−2
¯ωI
9.08 × 10−3
1.26 × 10−2
1.29 × 10−2
1.60 × 10−3
2.53 × 10−3
2.05 × 10−3
1.49 × 10−6
1.33 × 10−6
4.06 × 10−6
For m = −2, we compute free non-adiabatic g-modes, r-modes, and inertial modes of the planet model for ¯Ω = 0.1 and
tabulate their complex eigenfrequency in Table 2, where we have assumed τ∗ = 1day. Oscillation frequency ¯ω of the low
radial order g-modes in the envelope shows differences between prograde and retrograde modes for ¯Ω = 0.1 and ¯ω . 0.1. For
even parity r-modes of m = −2, the frequencies ¯ω tabulated in Table 2 are significantly different from the asymptotic value
2m ¯Ω/[l′(l′ + 1)], which is −1/3 for ¯Ω = 0.1. The inertial modes belonging to L ≡ l0 − m = 2 and 4 are tabulated in Table 2.
For the classification using l0 − m, see, e.g., Yoshida & Lee (2000), who computed inertial modes of isentropic polytropes to
tabulate κ0 = ω/Ω for different values of m and the polytropic index n, where κ0 was estimated in the limit of Ω → 0. The
ratio ω/Ω for the inertial modes in Table 2 in this paper is consistent with the value of κ0 computed for the m = 2 inertial
modes of the n = 1 polytrope, see Table 1 of Yoshida & Lee (2000). Note that for positive m, prograde (retrograde) modes
have negative (positive) κ0 for ¯Ω > 0. Since the inertial modes are confined in the convective core where non-adiabatic effects
are negligible, the imaginary part of the inertial mode frequency is much smaller than that of the g-modes and r-modes, which
are confined in the radiative envelope where non-adiabatic effects are very large.
Assuming only the thermal tides operate (i.e., j ∗ 6= 0 and ψ = 0), we compute the tidal torque N as a function of the
forcing frequency ¯ω for a fixed value of ¯Ω for τ∗ = 1day. The plots of N for ¯Ω = 0.05 and 0.1 are respectively given in Figures
3 and 4, where positive and negative ¯ω corresponds to prograde and retrograde forcing observed in the co-rotating frame of
the planet, and the red (black) lines represent positive (negative) parts of N . The left panels show the torque in the range
of ¯ω 6 1 and the right panels for ¯ω 6 0.2 as a magnification. The tidal torque only weakly depends on ¯ω for ¯ω & 0.2.
However, there appear broad and sharp peaks of N in the range of ¯ω . 0.2. Comparing the frequency ¯ωP at the peaks
with the natural frequency ¯ωR of the low frequency modes tabulated in Table 2, we find that the broad peaks are produced
when the forcing frequency is in resonance with the natural frequency of the g-modes and r-modes in the envelope, and that
the sharp peaks are produced by the resonance with the inertial modes in the core. The width of the peaks may reflect the
magnitude of ¯ωI of the modes in resonance with the forcing, that is, if the modes in resonance have ¯ωI ∼ ¯ωR, the peaks
will be broad, while if they have ¯ωI ≪ ¯ωR the peaks will be very sharp. Comparing the two cases of ¯Ω = 0.05 and 0.1, the
frequency ¯ωP of the broad peaks due to the g-modes does not significantly depend on ¯Ω, which is particularly the case for
the peaks on the prograde sides. It is also interesting to note that the frequency ¯ωP of the peaks due to the r-modes does
not show strong dependence on ¯Ω, which is because the frequency of r-modes propagating in a geometrically thin atmosphere
becomes insensitive to the rotation speed Ω for rapid rotation (see, e.g., Pedlosky 1986). The peak frequency ¯ωP due to the
inertial modes, however, linearly depends on the rotation frequency ¯Ω since the natural frequency ω ∝ Ω for inertial modes.
For example, for ¯Ω = 0.1, the sharp peaks located at the frequency ¯ω ≈ 0.055 and −0.11 respectively correspond to the
inertial modes with the ratio ω/Ω ≈ 0.55 and −1.1 belonging to l0 − m = 2 (see Yoshida & Lee 2000). For ¯Ω = 0.05, the
peak frequency ¯ωP is halved compared to that for ¯Ω = 0.1.
j=1 ρ′
lj (r)Y m
and ρ′(r, θ, φ) = Pjmax
The local tidal torque applied to a spherical surface is proportional to −ΦT Im(ρ′∗), where ΦT is given by equation (18),
lj (θ, φ) for m = −2 where jmax = 12 is used in this paper. In Figure 5 & 6, we show the
color-maps of −Im(ρ′) in the cos θ − log10 p plane, assuming φ = 0. Figure 5 is for the prograde and retrograde g1-modes and
the r1-mode at the forcing frequency tabulated in Table 2 for ¯Ω = 0.1 and Figure 6 for prograde and retrograde inertial modes
i2 at the forcing frequency ¯ω = 0.0557 and −0.11, respectively. The patterns are symmetric about the equator cos θ = 0. The
local tidal torque is confined into a geometrically very narrow region at the bottom of the radiative envelope and the direction
of the torque changes in this narrow layer, which could lead to a strong differential rotation there. The amplitudes of the
torque is confined in an equatorial region for the g-modes, and this confinement is stronger for the retrograde g1-mode. At
the forcing frequency of the r1-mode, the amplitude has two peaks as a function of cos θ and is small at the equator. At the
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
10
U. Lee, D. Murakami
Figure 5. Color maps of −Im(ρ′) for φ = 0 produced by semi-diurnal thermal tides, from left to right panels, at the forcing frequency
tabulated in Table 2 for the prograde and retrograde g1-modes and the r1-mode for ¯Ω = 0.1, where the pressure p is given in dyn/cm2
and the magnitudes of the torque are normalized by the maximum value.
Figure 6. Same as Figure 5 but at the forcing frequency ¯ω = 0.0557 and −0.11, respectively corresponding to the prograde and retrograde
inertial modes i2 in the core.
resonant forcing frequency for the i2 inertial modes in the core, the amplitude distribution for the retrograde inertial mode
is much more complicated than that for the prograde inertial mode, which has a similar distribution to that of the prograde
g1-mode.
Figure 7 shows the tidal torque N computed assuming j ∗ = 0 and ψ 6= 0 for τ∗ = 1day. There appears more sharp peaks
produced by resonance between the forcing and inertial modes in the core, compared to the case of pure thermal tides. The
broad peaks due to the g-mode resonance are pierced by such sharp peaks due to the inertial modes. Because the tidal potential
ΦT has substantial amplitudes in the convective core, the inertial modes in the core are more susceptible to the gravitational
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
35
34
33
32
31
30
29
28
27
-1
-0.5
0
0.5
1
Thermal Tides in Hot Jupiters
11
35
34
33
32
31
30
29
28
27
-0.2
-0.1
0
0.1
0.2
N
g
o
l
0
1
N
g
o
l
0
1
ω/σ
0
ω/σ
0
Figure 7. Same as Figure 4 but for j ∗ = 0 and ψ 6= 0.
tides than the thermal tides. We also find peaks due to the resonance with the envelope r-modes on the retrograde side. See
the Appendix C for a discussion about the alternative changing of the sign of N .
Instead of assuming the rotation rate ¯Ω takes a constant value, we let ¯Ω change as a function of ¯ω (or ¯ω changes as a
function of ¯Ω) for a given ¯Ωorb, that is, ¯Ω is given by
¯Ω = ¯Ωorb − ¯ω/2.
(50)
In Figure 8, we plot the tidal torque N as a function of the forcing period τtide = 2π/ω, where the left panel is for ¯Ωorb = 0.0537
and the right panel for ¯Ωorb = −0.0537. Since we assume τtide > 0, ¯Ω changes sign for ¯Ωorb = 0.0537 but it stays negative
for ¯Ωorb = −0.0537. Since prograde (retrograde) forcing corresponds to positive (negative) ν = 2Ω/ω, as τtide increases, the
forcing changes from retrograde to prograde for ¯Ωorb = 0.0537 and it is always retrograde for ¯Ωorb = −0.0537. We find the
gross properties of N as a function of τtide shown by the left panel of Figure 8 is similar to those computed by Auclaire-
Desrotour & Leconte (2018) using the traditional approximation, except that we have sharp resonance peaks due to inertial
modes in the convective core. The reasons for the difference may be partly because they used the traditional approximation,
with which inertial modes cannot be properly calculated, and partly because they assumed Γ1 = 1.4, for which the convective
core is not necessarily isentropic and propagation of inertial modes in the core may be suppressed. Assuming negative ¯Ωorb
(right panel), we can calculate retrograde forcing with long periods, with which the envelope r-modes are excited for τtide & 10
days.
4 CONCLUSION
We have computed the tidal torque due to thermal tides in rotating hot Jupiters, composed of a thin isothermal radiative
envelope and a nearly isentropic convective core. The thin envelope suffers the strong irradiation by the host star and
the periodic alternations of day and night sides on the planet produces semi-diurnal thermal tides. We have taken into
consideration radiative cooling in the envelope as the non-adiabatic energy dissipation mechanism. To represent the tidal
responses in rotating planets, we use series expansions in terms of spherical harmonic functions Y m
l (θ, φ) with different ls
for a given m. For fixed values of ¯Ω, we have computed the tidal torque as a function of the tidal forcing frequency ω for
both prograde and retrograde forcing, observed in the co-rotating frame of the planet. We find that at the forcing frequency
ω ∼pGM/R3, the tidal torque tends to synchronize the planet spin with the orbital motion, the direction of which is the
same as that by gravitational tides. At low frequency, the tidal forcing can be in resonance with low frequency modes such as
g-modes and r-modes in the envelope and inertial modes in the core and the resonance tends to enhance the tidal torques.
The sign of the tidal torque at the resonance peaks changes alternately as the mode that is in resonance with the forcing is
changed with ω. The tidal resonance with the g- and r-modes produces broad peaks of the torque and that with the inertial
modes sharp peaks as a function of the forcing frequency. The peak frequency ωP of the broad peaks by the g- and r-modes
is only weakly dependent on the spin frequency Ω and ωP of the sharp peaks is proportional to Ω.
We find a few differences between the results obtained in this paper and those by Auclair-Desrotour & Leconte (2018). One
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
12
U. Lee, D. Murakami
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
Ω<0
Ω>0
Ω
> 0
orb
N
g
o
l
0
1
33
32
31
30
29
28
27
26
25
Ω<0
Ω
< 0
orb
0.1
1
τ
tide
10
(day)
100
0.1
1
τ
tide
10
(day)
100
Figure 8. Tidal torque, given in erg, due to thermal tides for τ∗ = 1 day versus the tidal forcing period τtide = 2π/ω in days, where the
rotation speed Ω of the planet is given by Ω = Ωorb − π/τtide as a function of τtide for a given Ωorb, and we use ¯Ωorb = 0.0537 for the
left panel and ¯Ωorb = −0.0537 for the right panel. Here, the red lines and black lines respectively indicate positive and negative torque
N . The vertical dotted line in the left panel indicates the forcing period at which ¯Ω = 0. Note that ¯Ω < 0 ( ¯Ω > 0) corresponds to the
retrograde (prograde) forcing for τtide > 0.
of the differences may concern the core inertial modes of rotating Jovian planets. The traditional approximation employed by
Auclair-Desrotour & Leconte (2018) to represent the tidal responses in rotating planets cannot properly treat inertial modes
propagating isentropic regions. The Jovian models used in this paper and by Auclair-Desrotour & Leconte (2018) have the
convective core that has the structure of a polytrope of the index n = 1. For the core Auclair-Desrotour & Leconte (2018)
assumed Γ1 = 1.4 to avoid a nearly isentropic structure and hence suppressed core inertial modes. On the other hand, we
use series expansion in terms of spherical harmonic functions to represents tidal responses in rotating planets and assume
Γ1 = 2 to make the core nearly isentropic, which supports propagation of inertial modes. Because core inertial modes are
not necessarily susceptible to thermal tides prevailing in the radiative envelope, the difference between the present study and
Auclair-Desrotour & Leconte (2018) concerning the inertial modes may be considered as a minor difference, although at the
resonance peak with inertial modes the magnitude of tidal torque is significantly enhanced. Another difference between the
two analyses may concern the resonance with the envelope r-modes. Assuming Ωorb < 0, in this paper, we could compute
retrograde forcing with long periods and hence the tidal torques in resonance with the envelope r-modes. As shown by Figure
8, the behavior of tidal torques as a function of the forcing period τtide is different between prograde and retrograde forcing
with long periods, that is, as τtide increases the tidal torque on the prograde side stay positive to work for synchronization
but on the retrograde side it changes its sign alternatively.
We compute the rate of energy dissipation cause by thermal tides in the envelope where non-adiabatic effects are signifi-
cant. We define the normalized energy dissipation rate D as
D =
¯ω
2 Z R
0
Im Xl
δT ∗
l
T
δsl
cp ! ρT cpr3σ0
Leq
dr
r
,
(51)
where Leq ≡ 4πR2F∗ = 4πR2σSBT 4
∗ (R∗/r∗)2 (see Lee 2019). For the stellar parameters we use in this paper, we have
Leq ≈ 6 × 1029erg/s and hence Lextra/Leq ∼ 10−2 for Lextra ∼ 1027 − 5 × 1027erg/s, which is the magnitude of the extra heat
source needed to inflate the planets (see Baraffe et al 2003). As suggested by Figures 5 & 6, strong heating due to thermal
tides occurs in the bottom layers of the envelope. In Figure 9, we plot D as a function of the forcing frequency ¯ω for three
values of τ∗ for ¯Ω = 0.1. The dissipation rate has large values for ¯ω . 0.2, corresponding to the frequency range in which
tidal forcing can be in resonance with the low frequency modes in the envelope and inertial modes in the core. The magnitude
of D increases as τ∗ increases and it becomes D ∼ 10−3 for τ∗ = 10day, suggesting that non-adiabatic heating caused by
thermal tides at the bottom of the envelope can be a heating source for inflation of the planets if τ∗ is sufficiently long.
The results presented in this paper may depend on the expansion length jmax if the length is not long enough. We compute
for jmax = 20 the tidal torque as a function of the forcing frequency assuming j ∗ = 0 and ψ 6= 0, and the result is shown by
Figure 10. Comparing to Fig. 7, for which we assumed jmax = 12, we find that the tidal torque as a function of the forcing
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
Thermal Tides in Hot Jupiters
13
-2
10
-3
10
-4
10
D
-5
10
-6
10
-7
10
-8
10
-0.4
-0.2
0
0.2
0.4
ω/σ
0
Figure 9. Normalized energy dissipation rate D due to thermal tides versus forcing frequency ¯ω for τ∗ = 10day (dash-dotted line), 1day
(solid line), and 0.1day (dotted line) for ¯Ω = 0.1 where D is defined by equation (51).
N
g
o
l
0
1
35
34
33
32
31
30
29
28
27
N
g
o
l
0
1
35
34
33
32
31
30
29
28
27
-1
-0.5
0
0.5
1
-0.2
-0.1
0
0.1
0.2
ω/σ
0
ω/σ
0
Figure 10. Same as Figure 7 but for jmax = 20.
frequency ¯ω is almost the same between the cases of jmax = 12 and 20. We confirm that the length jmax = 12 is long enough
to produce reliable results.
2/ρ), which is the maximum value of ρ′
With an asymptotic treatment of waves, the local strength of nonlinearity of the waves could be discussed by using a
quantity ξrkr, where kr is the radial component of the wavenumber vector (e.g., Goodman & Dickson 1998). Here instead
we simply use the quantity max (ρ′
2/ρ in the interior of the planet, to consider the
validity of linear approximation employed in this paper. Here −Im(ρ′
2) is used to compute the tidal torque. In Fig. 11 we
plot max (ρ′
2/ρ) as a function of the forcing period τtide (day) for j ∗ 6= 0 and ψ = 0, which corresponds to Fig. 8, where for
a given value of ¯Ωorb the rotation speed Ω is given by Ω = Ωorb − π/τtide as a function of τtide. This figure shows that the
amplitudes of the tidal responses to pure thermal tides are less than 0.1 and stay in a linear regime for the parameters used
in this paper. So long as the amplitudes stay in a linear regime, the amplitudes are proportional to the external parameter
F∗, which depends on the luminosity of the host star and the distance between the host star and the planet. As suggested by
the figure, however, if the parameter F∗ increases by one or two order of magnitudes, the responses to the thermal tides enter
into a non-linear regime and we need non-linear treatment of the responses. In a nonlinear regime, the tidal responses excite
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
14
U. Lee, D. Murakami
0.1
0.08
0.06
0.04
0.02
0.1
0.08
0.06
0.04
0.02
2
)
ρ
/
'
ρ
(
x
a
m
2
)
ρ
/
'
ρ
(
x
a
m
Ω
>0
orb
0
0.1
1
τ
tide
10
(day)
Ω
<0
orb
100
0
0.1
1
τ
tide
10
(day)
100
Figure 11. max (cid:0)ρ′
speed Ω is given by Ω = Ωorb − π/τtide. This figure corresponds to Fig. 8.
2/ρ(cid:1) as a function of the forcing period τtide (day) for j ∗ 6= 0 and ψ = 0 and for τ∗ = 1day, where the rotation
many different oscillation modes by non-linear mode coupling, leading to a strong damping of the responses (e.g., Kumar &
Goodman 1996). Note that for pure gravitational tides (j ∗ = 0 and ψ 6= 0), we already have max (ρ′
2/ρ) ∼ 1, suggesting that
we need non-linear treatment of the responses, although the amplitudes are quite uncertain because we consider no dissipative
processes in the convective core in which the perturbing tidal potential ΦT has substantial amplitudes.
It is useful to make clear the relation between the methods of solutions used in this paper and by Auclair-Desrotour &
Leconte (2018) for thermal tides. To represent the tidal responses in rotating planets, we use series expansions in terms of
spherical harmonic functions Y m
l (θ, φ). The tidal torque on the planet, if we simply assume the tidal potential given by ΦT ∝
Y −2
2 in this equation is obtained by solving equations (45) and (46) and using equation
(10). Auclair-Desrotour & Leconte (2018), on the other hand, used series expansions of the responses in terms of the Hough
functions Θkm(θ; ν) defined in the traditional approximation (e.g., Lee & Saio 1997). Defining Θkm(θ, φ; ν) = fkmΘkmeimφ
, may be given by equation (49) and ρ′∗
2
where fkm is introduced so that the normalization D Θk′m ΘkmE = δk′k is satisfied and
hf gi =Z π
dθ sin θZ 2π
dφf ∗g,
0
0
we may have, assuming the functions Θkm form a complete set,
ylk Θkm,
Y m
l =Xk
ρ′(r, θ, φ) =Xk
k(r) Θkm.
ρ′
For the density perturbation ρ′ given by ρ′
2Y −2
2
, we obtain
ρ′
2
2 =(cid:10)Y −2
ρ′(cid:11) =Xk
ρ′
kDY −2
2
Θk,−2E =Xk
ρ′
ky∗
2k,
and for the tidal potential ΦT = ΦT (r)Y −2
ΦT,k Θk,−2
2 =Pk
(52)
(53)
(54)
(55)
ΦT,k =D Θk,−2ΦTE = ΦT (r)y2k.
Adding ΦT,k as an inhomogeneous forcing term, we compute the density perturbations ρ′
k in the traditional approximation.
The tidal torque due to equilibrium gravitational tides may be estimated as (e.g., Goldreich & Soter 1966)
Neq =
∗ R5
3GM 2
2a6
∗
1
Q
,
(56)
where Q is the tidal quality factor, representing the magnitude of the phase lag caused by energy dissipations that arise from
interaction between the tidal potential and fluid motion in the interior. The Q value for the interaction between the tidal
potential and the convective core is difficult to estimate since the fluid motion in the core is usually turbulent so that we need
properly treat effective viscosity for turbulence to estimate the amount of energy dissipations (see, e.g., Zahn 1977; Goldreich
& Nicholson 1977). For the parameters used in this paper, we have Neq = 1.6 × 1038/Q, which could be comparable to the
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
Thermal Tides in Hot Jupiters
15
torque due to the thermal tides calculated in this paper only for Q & 107, except for those at the peaks produced by resonance
with inertial modes. Probably, the magnitude Q & 107 is too large for Jovian planets (e.g., Goldreich & Nicholson 1977).
As discussed by Auclair-Desrotour & Leconte (2018), if we consider local timescales for the rotation rates to change in the
envelope and in the convective core, the two timescales can be comparable with each other for reasonable values of Q since
the moment of inertia of the thin envelope is much smaller that that of the convective core. If we assume certain formulae for
turbulent viscosity coefficient as done by Ogilvie & Lin (2004), we could estimate the tidal torque caused by both gravitational
and thermal perturbations although we have to solve the Navier Stokes equations for rotating planets, which will be one of
our future works.
APPENDIX A: DERIVATION OF THE OSCILLATION EQUATIONS
In this Appendix, we give a brief account of the derivation of the oscillation equations (30) to (34). The three components of
the perturbed equation of motion (9) are written as
−ρω2ξr − 2iωΩρξφ sin θ = −
∂p′
∂r
− ρ′ dΦ
dr
− ρ′ ∂ΦT
∂r
,
−ω2ρξθ − 2iωΩξφ cos θ = −
1
r
∂p′
∂θ
− ρ
1
r
∂ΦT
∂θ
,
−ω2ρξφ + 2iωΩρ (ξθ cos θ + ξr sin θ) = −
1
r sin θ
∂p′
∂φ
− ρ
1
r sin θ
∂ΦT
∂φ
.
(A1)
(A2)
(A3)
Substituting the expansions given by (25) to (28) into equation (A1), we find that the radial component of the equation of
motion (A1) reduces to
Xl (cid:0)−c1 ¯ω2Sl + 2c1 ¯ω ¯ΩmHl(cid:1) Y m
l + 2c1 ¯ω ¯ΩXl′
where
iTl′ sin θ
∂
∂θ
Y m
l′ =Xl
Pl
ρg
Y m
l
,
Pl = −ρgr
∂
∂r
Y2,l − ρg
d ln ρgr
d ln r
Y2,l + ρg
ΦT,l
gr
d ln ρ
d ln r
− ρg
ρ′
l
ρ
,
Y2,l =
p′
l
ρgr
+
ΦT,l
gr
.
(A4)
(A5)
(A6)
Similarly, using the θ and φ components of the perturbed equation of motion, sin−1 θ∂θ sin θ(eq. A2) + sin−1 θ∂φ(eq. A3),
which is the divergence of the horizontal displacement where ∂θ = ∂/∂θ and ∂φ = ∂/∂φ, gives
Xl (cid:0)c1 ¯ω2ΛlHl − 2c1 ¯ω ¯ΩmHl − 2mc1 ¯ω ¯ΩSl(cid:1) Y m
l −Xl′
2c1 ¯ω ¯Ω(cid:18)Λl′ iTl′ cos θ + iTl′ sin θ
∂
∂θ(cid:19) Y m
l′ =Xl
and sin−1 θ∂θ sin θ(eq. A3) − sin−1 θ∂φ(eq. A2), which corresponds to the radial component of ∇ × ξ, gives
ΛlY2,lY m
l
,
(A7)
The linearized continuity equation (10) may reduce to
l′ + 2c1 ¯ω ¯ΩXl (cid:18)ΛlHl cos θ + Hl sin θ
l +
Xl′ (cid:0)−c1 ¯ω2Λl′ iTl′ + 2c1 ¯ω ¯ΩmiTl′(cid:1) Y m
Xl (cid:18)ρ′
r3ρSl − ρΛlHl(cid:19) Y m
Xl (cid:20) δsl
and the entropy perturbation (23) to
ǫ′
l
T cp
iω + ωD
∂
∂r
1
r2
cp
−
+
1
Using the relations given by
l = 0,
sin θ
∂Y m
l
∂θ
= lJ m
l+1Y m
l+1 − (l + 1)J m
l Y m
l−1,
cos θY m
l = J m
l+1Y m
l+1 + J m
l Y m
l−1,
ωD
iω + ωD (cid:18)∇ad
δpl
p
+ ∇V Sl(cid:19)(cid:21) Y m
l = 0.
∂
∂θ
− 2Sl cos θ − Sl sin θ
∂
∂θ(cid:19) Y m
l = 0.
(A8)
(A9)
(A10)
(A11)
(A12)
where J m
l =p(l2 − m2)/(4l2 − 1) for l > m and J m
l = 0 otherwise, we rewrite each of the equations (A4), (A7), (A8), (A9),
l = 0. With the dependent variables as defined by equation (29), each set of the equations
· · · , jmax is written in the form as given by the oscillation equations (30) to (34). Note that equations
Alj = 0 for j = 1,
(A4), (A7), (A8), (A9), and (A10) correspond to equations (30), (33), (32), (31), and (34), respectively.
and (A10) into the form Pl AlY m
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
16
U. Lee, D. Murakami
1000
10
0.1
0.001
0
D
σ
/
ω
10-5
10-7
10-9
10-11
10-13
10-15
1012
1010
108
106
log
104
p
10
0
D
σ
/
ω
1000
10
0.1
0.001
-5
10
-7
10
-9
10
-11
10
-13
10
-15
10
100
1
0.01
0
0.2
0.4
0.6
0.8
1
r/R
Figure C1. ¯ωD as a function of log10 p (left panel) and of r/R (right panel) for τ∗ = 1day.
APPENDIX B: INNER BOUNDARY CONDITIONS
At the centre of the planet, the set of linear ordinary differential equations (45) and (46) can be formally written as
r
dz
dr
= Az,
z = (zj) = y1
Y 2 ! ,
(B1)
where A is the coefficient matrix for the differential equations and zj is for j = 1,
Assuming z ∝ rβ at the center and substituting into (B1) (see, e.g., Unno et al 1989), we obtain
· · · , 2jmax for the expansion length jmax.
(A − βI) z = 0,
(B2)
which gives 2jmax eigenvalues βj and eigenfunctions zj. Among the 2jmax eigenvalues, we pick up jmax eigenvalues βj that
satisfy the regularity condition given by Re(βj ) > −1 and the corresponding eigenfunctions zj . Using these eigenvalues and
eigenfunctions, we may represent the function z at the centre as
, Z = (z1, · · · , zjmax ),
(B3)
where Cj are arbitrary constants. Eliminating the terms Cjrβj , we obtain jmax linear relations between zj , which we use as
the inner boundary conditions.
z =
jmax
Xj=1
Cjrβj zj = Z
C1rβ1
...
Cjmax rβjmax
APPENDIX C: TIDAL TORQUE N AS A FUNCTION OF ω FOR ψ 6= 0
Since we take no account of dissipative processes in the convective core except for radiative damping associated with Newtonian
cooling, the results for the tidal torque obtained in this paper are not necessarily reliable, particularly for the case of ψ 6= 0.
The perturbing tidal potential ψ has substantial amplitudes in the core and hence the tidal responses to ψ in the core can be
strongly affected by dissipative processes there and so is the tidal torque N . The Newtonian cooling in the envelope considered
in this paper is controlled by the parameter ωD. Fig. C1 plots ¯ωD as a function of log10 p and r/R for τ∗ = 1day, and shows that
¯ωD increases by several orders of magnitudes within a geometrically thin layer near the bottom of the envelope from ¯ωD ∼ 10−5
at p ∼ pb to ¯ωD ∼ 0.1 at p ∼ p∗. For a given forcing frequency ¯ω, strong tidal torque is produced in the layer of ¯ω ∼ ¯ωD,
which occurs in this thin layer except in the limit of ¯ω → 0. As equation (49) indicates, the tidal response ρ′
2, particularly its
imaginary part, plays an essential role to determine the tidal torque. In Fig. C2, the tidal response −Im(ρ′
2/ρ) = Im(ρ′∗
2 /ρ)
and the cumulative tidal torque N (r) defined by
N (r) =r 3π
10
GM∗
∗ Z r
0
a3
drr4Im[ρ′∗
2 (r)]
(C1)
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
)
ρ
/
2
'
ρ
(
m
I
-
0.15
0.1
0.05
0
-0.05
-0.1
-0.15
1012
1010
108
106
log
104
p
10
Thermal Tides in Hot Jupiters
17
2E+32
1E+32
0
)
r
(
N
-1E+32
-2E+32
-3E+32
-4E+32
-5E+32
100
1
0.01
1012
1010
108
100
1
0.01
106
log
104
p
10
Figure C2. −Im(ρ′
j ∗ = 0 and ψ 6= 0. The solid and dotted curves respectively correspond to the forcing frequency ¯ω = 0.1 and 0.0812297.
2/ρ) (left panel) and N (r) (right panel) as a function of log10 p for ¯Ω = 0.1 and τ∗ = 1day, where we have assumed
are plotted for j ∗ = 0 and ψ 6= 0 for two forcing frequencies ¯ω = 0.1 and 0.0812297, which respectively correspond to positive
and negative N , where we use ¯Ω = 0.1 and τ∗ = 1day. As the figure indicates, significant changes of Im(ρ′∗
2 /ρ) and dN /d ln p
occur in the region of ¯ω ∼ ¯ωD and the tidal torque N is determined by the balance between positive and negative contributions
of dN /d ln p to N in the layer. The balance within this geometrically thin layer depends on the response ρ′
2 there and hence
on the forcing frequency ω. Note that we find similar behavior of −Im(ρ′
2/ρ) and N (r) also for the case of j ∗ 6= 0 and ψ = 0.
Because both forcing terms j ∗ and ψ in the perturbed entropy equation (34) obtained under the Newtonian
cooling approximation appear with the same factor 1/(iω + ωD), which is responsible for the deviation from
adiabatic perturbations, the behaviors of the thermal responses to j ∗ and ψ become similar in the envelope.
If we could correctly include the effects of dissipations in the convective core, the results for the tidal torque N would be
different from those computed in this paper, particularly when we consider tidal responses to ψ since the relation between
the entropy perturbation and the forcing ψ in the core will be different from the relation we use for the
envelope in this paper.
REFERENCES
Auclair-Desrotour P., Leconte J., 2018, A&A, 613, A45
Arras P., Socrates A., 2010, ApJ, 714, 1
Baraffe I., Chabrier G., Barman T.S., Allard F., Hauschildt P.H., 2003, A&A, 402, 701
Bodenheimer P., Lin D.N.C., Mardling R.A., 2001, ApJ, 548, 466
Clayton D.D., 1983, Principles of Stellar Evolution and Nucleosynthesis, The University of Chicago Press, Chicago
Fuller J., Lai D., 2013, MNRAS, 430, 274
Goldreich P., Nicholson P.D., 1977, Icarus, 30, 301
Goldreich P., Soter S., 1966, Icarus, 5, 375
Goodman J., Dickson E.S., 1998, ApJ, 507, 938
Greenspan H.P., 1969, The Theory of Rotating Fluids, Cambridge University Press, Cambridge
Iro N., B´ezard B., Guillot T., 2005, A&A, 436, 719
Ivanov P.B., Papaloizou J.C.B., 2007, MNRAS, 376, 682
Jermyn A.D., Tout C.A., Ogilvie G.I., 2017, MNRAS, 469, 1768
Kumar P., Goodman J., 1996, ApJ, 466, 946
Lai D., 1997, ApJ, 490, 847
Lee U., 2019, MNRAS, 484, 5845
Lee U., Saio H., 1986, MNRAS, 221, 365
Lee U., Saio H., 1987, MNRAS, 224, 513
Lee U., Saio H., 1997, ApJ, 491, 839
Mihalas D., Weibel-Mihalas B., 1999, Foundations of Radiation Hydrodynamics, Dover Publishing, New York
Ogilvie G.I., 2014, Annu. Rev. Astron. Astrophys., 52, 171
Ogilvie G.I., Lin N.D.C., 2004, ApJ, 610, 477
Papaloizou J., Pringle J.E., 1978, MNRAS, 182, 423
Press W.H., Teukolsky S.A., 1977, ApJ, 213, 183
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
18
U. Lee, D. Murakami
Savonije G.J., Papaloizou J.C.B., 1984, MNRAS, 207, 685
Savonije G.J., Papaloizou J.C.B., 1997, MNRAS, 291, 633
Stevenson D.J., Geophys. Astrophy. Fluid. Dynamics, 1979, 12, 139
Stevenson D.J., Salpeter E.E., 1977a, ApJS, 35, 221
Stevenson D.J., Salpeter E.E., 1977b, ApJS, 35, 239
Turner J.S., 1979, Buoyancy Effects in Fluids, Cambridge University Press, Cambridge
Unno W., Osaki Y., Ando H., Saio H., Shibahashi H., 1989, Nonradial Oscillations of Stars, 2nd ed., University of Tokyo Press, Tokyo
Witte M.G., Savonije G.J., 2002, A&A, 386, 222
Yoshida S., Lee U., 2000, ApJ, 529, 997
Zahn J.P., 1977, A&A, 57, 383
c(cid:13) 2019 RAS, MNRAS 000, 000 -- 000
|
1512.00134 | 3 | 1512 | 2016-05-12T07:32:30 | The First Neptune Analog or Super-Earth with Neptune-like Orbit: MOA-2013-BLG-605Lb | [
"astro-ph.EP"
] | We present the discovery of the first Neptune analog exoplanet or super-Earth with Neptune-like orbit, MOA-2013-BLG-605Lb. This planet has a mass similar to that of Neptune or a super-Earth and it orbits at $9\sim 14$ times the expected position of the snow-line, $a_{\rm snow}$, which is similar to Neptune's separation of $ 11\,a_{\rm snow}$ from the Sun. The planet/host-star mass ratio is $q=(3.6\pm0.7)\times 10^{-4}$ and the projected separation normalized by the Einstein radius is $s=2.39\pm0.05$. There are three degenerate physical solutions and two of these are due to a new type of degeneracy in the microlensing parallax parameters, which we designate "the wide degeneracy". The three models have (i) a Neptune-mass planet with a mass of $M_{\rm p}=21_{-7}^{+6} M_{Earth}$ orbiting a low-mass M-dwarf with a mass of $M_{\rm h}=0.19_{-0.06}^{+0.05} M_\odot$, (ii) a mini-Neptune with $M_{\rm p}= 7.9_{-1.2}^{+1.8} M_{Earth}$ orbiting a brown dwarf host with $M_{\rm h}=0.068_{-0.011}^{+0.019} M_\odot$ and (iii) a super-Earth with $M_{\rm p}= 3.2_{-0.3}^{+0.5} M_{Earth}$ orbiting a low-mass brown dwarf host with $M_{\rm h}=0.025_{-0.004}^{+0.005} M_\odot$ which is slightly favored. The 3-D planet-host separations are 4.6$_{-1.2}^{+4.7}$ AU, 2.1$_{-0.2}^{+1.0}$ AU and 0.94$_{-0.02}^{+0.67}$ AU, which are $8.9_{-1.4}^{+10.5}$, $12_{-1}^{+7}$ or $14_{-1}^{+11}$ times larger than $a_{\rm snow}$ for these models, respectively. The Keck AO observation confirm that the lens is faint. This discovery suggests that low-mass planets with Neptune-like orbit are common. So processes similar to the one that formed Neptune in our own Solar System or cold super-Earth may be common in other solar systems. | astro-ph.EP | astro-ph | The First Neptune Analog or Super-Earth with Neptune-like
Orbit: MOA-2013-BLG-605Lb
T. Sumi1,2, A. Udalski3,4, D.P. Bennett5,6,2, A. Gould7, R. Poleski3,4,7, I.A. Bond8,2,
J.Skowron3,4, N. Rattenbury9,2, R. W. Pogge7, T. Bensby10, J.P. Beaulieu11 ,
J.B. Marquette11 , V. Batista11 , S. Brillant12
and
F. Abe13, Y. Asakura13, A. Bhattacharya5, M. Donachie9, M. Freeman9, A. Fukui14,
Y. Hirao1, Y. Itow13, N. Koshimoto1, M.C.A. Li9, C.H. Ling8, K. Masuda13,
Y. Matsubara13, Y. Muraki13, M. Nagakane1, K. Ohnishi15, H. Oyokawa13, To. Saito16,
A. Sharan9, D.J. Sullivan17, D. Suzuki5, P.,J. Tristram18, A. Yonehara19,
M.K. Szyma´nski3, K. Ulaczyk3, S. Koz lowski 3, L. Wyrzykowski3, M. Kubiak3, P.
(The MOA Collaboration)
Pietrukowicz3, G. Pietrzy´nski3, I. Soszy´nski3,
(The OGLE Collaboration)
and
C. Han20, Y.-K. Jung20, I.-G Shin20, C-U. Lee21
6
1
0
2
y
a
M
2
1
.
]
P
E
h
p
-
o
r
t
s
a
[
3
v
4
3
1
0
0
.
2
1
5
1
:
v
i
X
r
a
-- 2 --
ABSTRACT
We present the discovery of the first Neptune analog exoplanet or super-
Earth with Neptune-like orbit, MOA-2013-BLG-605Lb. This planet has a mass
1Department of Earth and Space Science, Graduate School of Science, Osaka University, Toyonaka, Osaka
560-0043, Japan,
e-mail: [email protected]
2Microlensing Observations in Astrophysics (MOA)
3Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa,Poland; udalski, msz, mk, pietrzyn,
soszynsk, szewczyk, [email protected]
4Optical Gravitational Lens Experiment (OGLE)
5Department of Physics, University of Notre Dame, Notre Dame, IN 46556, USA; [email protected]
6Laboratory for Exoplanets and Stellar Astrophysics, NASA/Goddard Space Flight Center, Greenbelt,
MD 20771, USA
7Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA
8Institute of Information and Mathematical Sciences, Massey University, Private Bag 102-904, North
Shore Mail Centre, Auckland, New Zealand; i.a.bond,c.h.ling,[email protected]
9Department of Physics, University of Auckland, Private Bag 92019, Auckland, New Zealand;
n.rattenbury,mli351,[email protected]; asha583,[email protected]
10Lund Observatory, Department of Astronomy and Theoretical physics, Box 43, SE-221 00 Lund, Sweden
11UPMC-CNRS, UMR7095, Institut d'Astrophysique de Paris, F-75014 Paris, France
12European Southern Observatory (ESO), Karl-Schwarzschildst. 2, D-85748 Garching, Germany
13 Institute for Space-Earth Environmental Research, Nagoya University, Nagoya 464-8601, Japan;
abe,itow,kmasuda,[email protected]
14Okayama Astrophysical Observatory, National Astronomical Observatory of Japan, 3037-5 Honjo, Kamo-
gata, Asakuchi, Okayama 719-0232, Japan
15Nagano National College of Technology, Nagano 381-8550, Japan
16Tokyo Metropolitan College of Aeronautics, Tokyo 116-8523, Japan
17School of Chemical and Physical Sciences, Victoria University, Wellington, New Zealand
18Mt. John University Observatory, P.O. Box 56, Lake Tekapo 8770, New Zealand
19Department of Physics, Faculty of Science, Kyoto Sangyo University, 603-8555 Kyoto, Japan
20Department of Physics, Institute for Astrophysics, Chungbuk National University, 371-763 Cheongju,
Korea
21Korea Astronomy and Space Science Institute, 305-348 Daejeon, Korea
-- 3 --
similar to that of Neptune or a super-Earth and it orbits at 9 ∼ 14 times the
expected position of the snow-line, asnow, which is similar to Neptune's separation
of 11 asnow from the Sun. The planet/host-star mass ratio is q = (3.6±0.7)×10−4
and the projected separation normalized by the Einstein radius is s = 2.39±0.05.
There are three degenerate physical solutions and two of these are due to a new
type of degeneracy in the microlensing parallax parameters, which we designate
"the wide degeneracy". The three models have (i) a Neptune-mass planet with
a mass of Mp = 21+6
−7M⊕ orbiting a low-mass M-dwarf with a mass of Mh =
0.19+0.05
−1.2M⊕ orbiting a brown dwarf
host with Mh = 0.068+0.019
−0.3M⊕
orbiting a low-mass brown dwarf host with Mh = 0.025+0.005
−0.004M⊙ which is slightly
favored. The 3-D planet-host separations are 4.6+4.7
−0.2 AU and 0.94+0.67
−0.02
AU, which are 8.9+10.5
times larger than asnow for these models,
respectively. The Keck AO observation confirm that the lens is faint. This
discovery suggests that low-mass planets with Neptune-like orbit are common.
So processes similar to the one that formed Neptune in our own Solar System or
cold super-Earth may be common in other solar systems.
−0.011M⊙ and (iii) a super-Earth with Mp = 3.2+0.5
−0.06M⊙, (ii) a mini-Neptune with Mp = 7.9+1.8
−1.4 , 12+7
−1 or 14+11
−1
−1.2 AU, 2.1+1.0
Subject headings: gravitational lensing: micro -- planetary systems
1.
Introduction
The formation of the ice giants Uranus and Neptune is not well understood.
In the
favored core accretion theory, the gas giant planets like Jupiter and Saturn are believed
to form through the accumulation of small icy planetesimals into solid cores of about
5-15 M⊕ in the region beyond the snow-line at asnow ≈ 2.7(Mh/M⊙) (Ida & Lin 2004;
Laughlin, Bodenheimer & Adams 2004; Kennedy, Kenyon & Bromley 2006), where the pro-
toplanetary disk is cold enough for ices (especially water-ice) to condense. However, such
a scenario is unable to form smaller ice giants like Uranus and Neptune at their current
orbital positions, due to the low density of planetesimals and slow evolution in these orbits
(Pollack et al. 1996). One idea is that Uranus and Neptune formed in the Jupiter-Saturn re-
gion between ∼ 5 and ∼17 AU, then migrated outwards to the current position (Fernandez
1984; Thommes, Duncan & Levison 1999; Helled & Bodenheimer 2013).
The formation of super-Earth exoplanets with Neptune-like orbit are even less under-
stood. This is partly because we do not even know if they exist contrary to the case of
Neptune. They are also not expected by the standard core accretion theory due to the same
reason as that of ice giants mentioned above (Ida & Lin 2004). Their formation may be
-- 4 --
similar to or related to the formation of Neptune-like ice giants.
The distribution of such cold ice-giant planets and super-Earth in other solar systems
is important for understanding the formation of our own cold ice giants. Also, in our own
Solar System, the distribution of Kuiper Belt objects (KBOs) is dominated by gravitational
interactions with Neptune. Since KBOs hold large amounts of water and other volatiles
needed for life, it could be that exo-neptunes play an important role in the development
of life in some exoplanetary systems, whether or not they play this role in our own Solar
System.
In the 20 years since the first exoplanet discovery (Mayor & Queloz 1995), there have
been repeated discoveries of planets that are quite different from those in our own So-
lar System. However, the detection of planets similar to those in our own Solar System
has been more difficult. Only Jupiter analogs have been detected orbiting solar type stars
(Wittenmyer et al. 2014), while Jupiter/Saturn (Gaudi et al. 2008; Bennett et al. 2010) and
Venus/Earth analogs (Burke et al. 2014; Quintana et al. 2014) have been found orbiting
low-mass stars. Very cold, low-mass planets have yet to be explored (see the distribution of
known exoplanets as of 2015 Oct. 6 1 in Figure 1). Cold ice-giants like Uranus and Neptune
are very difficult to detect with the radial velocity and transit methods owing to their long
orbital periods (80-160 years), low orbital velocities and low transit probabilities. It is even
more difficult to detect a super-Earth in such wide orbit. The direct imaging method can
detect wide-orbit planets if they are self-luminous, but otherwise, they will be far too faint
to detect, especially if they are as small as Neptune and super-Earth.
Recently, low-mass stars (i.e. M-dwarfs) have attracted more interest in exoplanet search
programs because of their high detectability of habitable or cold low-mass planets.
Kepler 's 150,000 targets contain about 3000 red dwarfs and more than a hundred plan-
etary systems have been found orbiting these stars (Morton & Swift 2013). These results
show that smaller planets are more common than larger planets around M-dwarfs, and plan-
ets with radii of ∼1.25 R⊕ are the most common planets in these systems. Dressing & Charbonneau
(2013) estimated an occurrence rate of ∼ 0.5 habitable zone Earth size planets per M-dwarf,
and Quintana et al. (2014) found an Earth-radius habitable planet around a ∼0.5 M⊙ M-
dwarf. That smaller planets are more common than larger planets around M-dwarfs may be
related to the fact that only small mass proto-planetary disks have been found around such
low mass stars (Kennedy & Kenyon 2008). The TRENDS high-contrast imaging survey, in
combination with radial velocity measurements, indicates that 6.5%± 3.0% of M-dwarf stars
host one or more massive companions with 1 < m/MJ < 13 and 0 < a < 20 AU, however
1http://exoplanet.eu
-- 5 --
this survey is not sensitive to cold ice planets (Montet et al. 2014; Clanton & Gaudi 2014).
The gravitational microlensing method is also sensitive to planets around M-dwarfs and
even brown dwarfs because it does not rely on the light from the host stars. Microlens-
ing relies upon random alignments between background source stars and foreground lens
star+planet systems, and more massive lens stars are only favored by the factor √M while
smaller masses have shorter timescales which can also bias against detection. So M-dwarf
lens stars dominate microlensing events. Contrary to the other methods, microlensing is sen-
sitive to low-mass planets down to an Earth-mass (Bennett & Rhie 1996) orbiting beyond
the snow-line, as shown in Fig. 1. Microlensing is therefore complementary to the other
planet detection techniques. Statistical analyses of microlensing samples indicate that the
planet abundance beyond the snow line is about a factor ∼ 7 larger than the abundance of
close-in planets. Neptune mass planets are more abundant than gas giants around M-dwarfs,
and one or more planets per star in total are predicted just beyond the snow-line (Sumi et al.
2010; Gould et al. 2010; Cassan et al. 2012).
In about half of the planetary systems found by microlensing, the mass of the host
and planets and their projected separation have been measured by microlensing paral-
lax in combination with the finite source effect (Bennett et al. 2008; Gaudi et al. 2008;
Muraki et al. 2011; Kains et al. 2013; Tsapras et al. 2014; Udalski et al. 2015) and/or direct
detection of the lens flux by high resolution imaging by adaptive optics (AO) (Bennett et al.
2010; Kubas et al. 2012; Batista et al. 2014, 2015) or the Hubble Space Telescope (HST)
(Bennett et al. 2006, 2015; Dong et al. 2009b). The probability distribution of physical
mass and separations of other events have been estimated using a Bayesian analysis as-
suming a Galactic model. Among the planetary systems with mass measurements, two
less than 0.2M⊙ and each system has a planetary
of them have very low mass hosts,
mass ratio q < 0.01. These two systems are MOA-2007-BLG-192L (Mh = 0.084+0.015
−0.012M⊙,
Mp = 3.2+5.2
−1.8M⊕) (Bennett et al. 2008; Kubas et al. 2012) and MOA-2010-BLG-328L (Mh =
0.11 ± 0.01M⊙, Mp = 9.2 ± 2.2M⊕) (Furusawa et al. 2013). Neptune analog planets and
super-Earth with Neptune-like orbit are still difficult to detect even by microlensing.
Recently, Poleski et al. (2014) found a planet in a Uranus-like orbit with mass of ∼4
MUranus at ∼18 AU around ∼0.7 M⊙ star. This is ∼ 9 times the snow-line of the host. While
their mass estimates are based on a Bayesian analysis and have large uncertainties, their
detection demonstrated the ability to detect planets in these orbits with microlensing.
In this paper, we present the detection and the mass measurement of the first Neptune
analog MOA-2013-BLG-605Lb via microlensing. We detected the microlensing parallax ef-
fect which yield the mass measurement of the lens system in combination with the finite
source effect.
-- 6 --
Microlensing parallax can be measured when one observes an event simultaneously from
two different locations, either with a telescope on Earth and a space telescope, (Refsdal 1966;
Udalski et al. 2015) or with two ground-based telescopes, referred to as terrestrial parallax
(Gould et al. 2009). It is known that there is a four-fold degeneracy in these parallax mea-
surements, (Refsdal 1966; Gould 1994). Two elements of this four-fold degeneracy correspond
to two different magnitudes of the measured parallax. As a result, the physical parameters of
the lens differs between these two degenerate solutions. The other two degenerate solutions
in the four-fold degeneracy just arise from a symmetry in the lensing geometry. The physi-
cal parameters of the lens are the same between these two degenerate solutions, except the
projected velocities which can be used to distinguish among solutions (Calchi Novati et al.
2015a). Most commonly, parallax measurements have been made by observing an event from
an accelerated observatory; specifically from ground-based observations of an event which
is long enough for Earth to move significantly in its orbit around the Sun. This is referred
to as orbital parallax (Gould 1992). There is also an analogous four-fold discrete degener-
acy for orbital parallax, termed the "jerk parallax" degeneracy and their mirror solutions
(Gould 2004; Park et al. 2004). For the binary lens case, there is an approximate degeneracy
in the parallax parameters, known as the "ecliptic degeneracy" (Skowron et al. 2011). In
this work on event MOA-2013-BLG-605, we report a new type of degeneracy in parallax
model solutions, which is specific to widely separated binary lenses. The details of this new
degeneracy are presented in section § 4.
We describe the observations of, and photometric data for, event MOA-2013-BLG-605
in sections § 2 and § 3. The light curve modeling is described in section § 4. In section § 5
and § 6 we present the physical parameters of the lens system and constraints by the Keck
AO observation. We discuss, in section § 7, the manner in which we might measure the lens
mass in the future and we present an overall discussion and our conclusions in section § 8.
2. Observation
The Microlensing Observations in Astrophysics (MOA; Bond et al. 2001; Sumi et al.
2003) collaboration carries out a microlensing survey toward the Galactic bulge from the
Mt. John University Observatory in New Zealand. The MOA-II survey (Sumi et al. 2011)
is a very high cadence photometric survey of the Galactic bulge with the 1.8 m MOA-II
telescope equipped with a 2.2 deg2 field-of-view (FOV) CCD camera. The 2013 MOA-II
observing strategy called for the 6 fields (∼ 13 deg2) with the highest lensing rate to be
observed with a 15 minute cadence, while the next 6 best fields were observed with a 47
minute cadence, and 8 additional fields were observed with a 95 minutes cadence. Most
-- 7 --
MOA-II observations use the custom MOA-red wide band filter, which corresponds to the
sum of the standard Cousins R and I-bands. MOA-II issues ∼ 600 alerts of microlensing
events in real time each year.2
The Optical Gravitational Lensing Experiment (OGLE; Udalski, Szyma´nski and Szyma´nski
2015) also conducts a microlensing survey toward the Galactic bulge with the 1.3 m Warsaw
telescope at the Las Campanas Observatory in Chile. The fourth phase of OGLE, OGLE-
IV started its high cadence survey observations in 2010 with a 1.4 deg2 FOV mosaic CCD
camera. OGLE observes bulge fields with cadences ranging from one observation every 20
minutes for 3 central fields to less than one observation every night for the outer bulge fields.
Most observations are taken in the standard Kron-Cousin I-band with occasional observa-
tions in the Johnson V -band. OGLE-IV issues ∼ 2000 microlensing event alerts in real time
each year.3
The microlensing event MOA-2013-BLG-605 was discovered at (α, δ)(2000) = (17:58:42.85,
-29:23:53.66) [(l, b) = (1.0583◦, -2.695◦)], in MOA field gb9, which is monitored every 15 min,
and it was announced by the MOA Alert System on 2013 Aug 30 (HJD′ ≡HJD−2450000 ∼
6535). Figure 2 shows the light curve. At the time of its discovery, MOA recognized this
event as a possible free-floating planet candidate (Sumi et al. 2011) as the best fit single lens
light curve had an Einstein radius crossing time of tE = 0.73 ± 0.10 days (See Figure 2).
Nearly four weeks later, the OGLE Early Warning System (EWS) (Udalski 2003) detected
this event being magnified again with longer timescale due to the lensing effect of the host
star. The OGLE EWS system announced this event as OGLE-2013-BLG-1835 on 2013 Sep
25 (HJD′ ∼ 6560), as shown in the top panel of Figure 2. The initial short magnification by
the planet at HJD′ ∼ 6535 was confined by the OGLE survey data. Actually, it should have
triggered the OGLE discovery alert but due to unfortunate deeply hidden bug in the EWS
software this did not happen. The later magnification by the host was observed by MOA,
as well as OGLE.
Follow-up observations of the stellar part of the light curve in the V , I and H-bands were
obtained by the µFUN collaboration using the SMARTS-CTIO 1.3 m telescope. These data
were taken mainly to extract the source color. We use the average of these CTIO and OGLE
V − I color measurements. CTIO H-band measurements are used to drive H-band source
magnitude, which is very important for comparison to the AO observations (see Section 6).
2https://it019909.massey.ac.nz/moa/
3http://ogle.astrouw.edu.pl/ogle4/ews/ews.html
-- 8 --
3. Data Reduction
The MOA images were reduced with MOA's implementation (Bond et al. 2001) of the
difference image analysis (DIA) method (Tomany & Crotts 1996; Alard & Lupton 1998;
Alard 2000).
In the MOA photometry, we found that there were systematic errors that
correlate with the seeing and airmass, as well as the motion, due to differential refraction,
of a nearby, possibly unresolved star. There is also a potential systematic error due to the
relative proper motion of a nearby star or stars, which we model as linear function in time.
We ran a detrending code to measure these effects in the 2011, 2012 and 2014 data, and we
removed these trends with additive corrections to the full 2011-2014 data set. (The MOA
data from 2006-2010 indicate no significant photometric variations, but they are not included
in the light curve analysis.) This detrending procedure improved the fit χ2 by ∆χ2 = 0.073
per data point in the baseline, so it has reduced the systematic photometry errors signifi-
cantly. This investigation of the systematics is necessary to have confidence in the modeling
of the light curve with high order effects in the following section.
The OGLE data were reduced with the OGLE DIA (Wo´zniak 2000) photometry pipeline
(Udalski, Szyma´nski and Szyma´nski 2015). In this event, the center of the magnified source
star is slightly shifted from the center of the apparent star identified in the reference image,
due to blending with one or more unresolved stars. So the OGLE data have been re-reduced
with a centroid based difference images, just as the MOA pipeline does (Bond et al. 2001).
The number of data points used for the light curve modeling are 9675, 5514 and 64
for MOA-Red, OGLE-I and OGLE-V passbands, respectively. The photometric errorbars
provided by the photometry codes give approximate estimates of the absolute photometric
uncertainty of each measurement, and we regard them as an accurate representation of the
relative uncertainty for each measurement. This is adequate for determining the best light
curve model, but in order to determine the uncertainties on the model parameters, it is
important to have more accurate error bars. We accomplish this with the method presented
in Yee et al. (2012). We rescale the errors using the formula, σ′
min, where σi
and σ′
i are original and renormalized errorbars in magnitudes. The parameters k and emin
are selected so that the cumulative χ2 distribution sorted by the magnification of the best
model is a straight line of slope 1 and χ2/dof∼ 1. This procedure yields k = 1.092313
and emin = 0.012662 for MOA-Red, k = 1.387059 and emin = 0.010938 for OGLE-I and
k = 1.571492 and emin = 0.0 for OGLE-V . Note that the changes of the final best fit model
due to this error renormalization are negligible.
i = kpσ2
i + e2
CTIO data were reduced by DoPHOT (Schechter, Mateo & Saha 1993), the point spread
function (PSF)-fitting routine. The number of data points in the CTIO-I, V and H pass-
bands is 15, 15 and 149, respectively. Their error bars are not rescaled, i.e., k = 1.0 and
-- 9 --
emin = 0.0. These CTIO data are not used for light curve modeling, but used for obtaining
the source color in (V −I) and (I−H) in a model-independent way from the linear regression
of these light curves by following Dong et al. (2009a) and Calchi Novati et al. (2015b).
Details of the datasets are summarized in Table 1.
4. Light curve modeling
We search for the best fit models of the standard (static), the parallax, the parallax
with the linear orbital motion of the planet, the keplerian orbital motion, with the keplerian
prior, the Galactic kinematic constraint and the Galactic density prior using Markov Chain
Monte Carlos (MCMC) (Verde et al. 2003). The best fit models are shown in Table 2-7 and
their physical parameters are in Tables 8-13, respectively (see Section 5).
4.1. Standard (static) model
In a point-source point-lens (PSPL) microlensing model, there are three parameters, the
time of peak magnification t0, the Einstein radius crossing time tE, and the minimum impact
parameter u0. The standard binary lens model has four more parameters, the planet-host
star mass ratio q, the projected separation normalized by Einstein Radius s, the angle of
the source trajectory relative to the binary lens axis α, and the ratio of the angular source
radius to the angler Einstein radius ρ = θ∗/θE. ρ can only be measured for events that show
finite source effects. The measurement of ρ is important because it allows us to determine
the angular Einstein radius θE = θ∗/ρ since the angular source radius, θ∗, can be estimated
from its color and extinction-corrected apparent magnitude (Kervella et al. 2004).
We use linear limb-darkening models for the source star using the coefficients, u =
0.5863, 0.7585 and 0.6327 for the I, V and MOA-Red bands, respectively (Claret 2000).
The MOA-Red value is the mean of the R and I-band values. These values were selected
from Claret (2000) for a K2 type source star with T = 5000 K, logg = 4.0 and log[M/H] = 0,
based on the extinction corrected, best fit source V − I color and brightness (see Section 5).
Initially, the global grid search of the best fit model was conducted with 9,680 fixed grid
points across a wide range of three parameters, −4.0 < log q < 0.4, −0.5 < log s < 0.6 and
0 < α < 2π, with all other parameters being free. Then the most likely models were refined,
allowing all parameters to vary. Using this robust search methodology, we avoid missing any
local minimum solutions across the wide range of parameter space. We found that only the
model with a wide separation (s > 1) reproduces the observed light curve data. The model
-- 10 --
corresponds to the source crossing a planetary caustic. Planetary caustics can form far from
the primary and any source star that crosses or passes close to such a distant a planetary
caustic will impose an signal far from the main microlensing peak. Furthermore, the shape of
planetary caustics differ significantly between wide (s > 1) and close (s < 1) configurations
in contrast to the close/wide degeneracy for an event crossing a central caustic near the
primary. The best fit standard model parameters are shown in Table 2. The mass ratio of
q ∼ 3 × 10−4 and separation of s ∼ 2.3 indicate that the companion is a relatively low mass
planet at wide separation.
The single lens model with a binary source was ruled out as follows. We extracted 88
data points around the planetary anomaly within 6533.0 <HJD′ < 6536.6 after subtracting
the flux contribution from the best fit single lens model of the primary peak which is fitted
without data around the anomaly. We fitted this extracted light curve by the single lens
model with a finite source effect. The best fit χ2 is ∼31 larger than the χ2 contributions
from the same data points by the planetary models with parallax and orbital motion Kpk
(see §4.4.3). Furthermore, the best fit event timescale of tE = 2.3 days is much smaller than
tE ∼ 20 days of the main peak, while they should be same if a single lens caused both two
magnifications.
4.2. Parallax model with a New Type of Degeneracy
There are higher order effects that require additional parameters. The orbital motion
of the Earth can cause the apparent lens-source relative motion to deviate from a constant
velocity. This effect is known as the microlensing parallax effect (Gould 1992; Alcock et al.
1995; Smith, Mao & Wo´zniak 2002), and it can be described by the microlensing parallax
vector πE = (πE,N, πE,E). The direction of πE is the direction of the lens-source relative
motion projected on the sky (geocentric proper motion at a fixed time), and the amplitude of
the microlensing parallax vector, πE = AU/rE, is the inverse of the Einstein radius, projected
to the observer plane. Because the Galactic bulge is close to the ecliptic plane, there is an
approximate degeneracy in the parallax parameters, known as the "ecliptic degeneracy,"
where models with similar parameters but with (u0, α, πE,N) = −(u0, α, πE,N) produce nearly
indistinguishable light curves. This corresponds to a reflection of the lens plane with respect
to the geometry of Earth's orbit, (Smith, Mao & Paczy´nski 2003; Skowron et al. 2011).
We found the four degenerate parallax models as shown in Table 2. The light curves of
these four models are almost identical to the one shown in Figure 2. The caustics, critical
curves and source trajectory of these models are shown in Figure 3. The "P" scripts indicate
models with microlensing parallax. The "+" and "−" subscripts refer to two different 2-
-- 11 --
fold degeneracies in the parallax models. The first "±" subscript refers to the sign of the u0
parameter, and refers to the "ecliptic degeneracy" mentioned above, but the second "±" sub-
script refers to a new parallax degeneracy, the wide degeneracy, that is particular to events
like this, with a wide separation planet detected through a crossing of the planetary caustic.
The light curve measurements indicate the angle, α(tpcc), between the source trajectory and
lens axis at the time of the planetary caustic crossing, tpcc. Due to the reflection symmetry
of the lens system, the light curve constrains α(tpcc) up to a reflection symmetry, as shown in
Figure 3. If there were no microlensing parallax, we could use α(tpcc) to predict the closest
approach of the source to the center-of-mass, u0, and therefore the peak magnification of the
the stellar part of the microlensing light curve. But, when the microlensing parallax effect is
included, the angle α can vary in time, so that α(tpcc) 6= α(t0). For a wide-separation plan-
etary event, like MOA-2013-BLG-605, the light curve basically constrains the microlensing
parallax through the three parameters, α(tpcc), u0, and t0, which is essentially the time of the
stellar peak magnification (cf. An & Gould 2001). As a result, the configurations shown in
the upper and lower panels of Figure 3 yield nearly identical light curves as shown in Figure
2, even though the source passes through in between the two masses in the upper panels
and below or above the masses in the bottom two panels. The lower panels imply a larger
curvature of the source trajectory, and therefore, a larger microlensing parallax signal. (Note
that the model parameter α0 and s0 given in Table 2-7 are the α and s values at a fixed time
tfix = 6573.045, following the convention of Geocentric microlensing parallax parameters.)
Fig. 4 shows the ∆χ2 distribution of the parallax parameters from the best fit MCMC
models. The best fit values are compared to that of other models in Figure 6.
In late
September, the Earth's acceleration is in the East-West (E-W) direction, so for a typical
event, we would expect a better constraint on parallax in the E-W direction, i.e, a smaller
error for πE,E. However, in this case, the planetary signal plays a big role in the parallax
signal. The angle and timing of caustic entry for a given u0 value -- which is constrained by
the main peak corresponding to the host star -- constrain the parallax parameters.
Figure 5 shows the difference in the cumulative χ2 values between the standard and the
parallax models as a function of time. We can see that most of the parallax signals come
from around the planetary signal in both MOA-Red and OGLE-I as expected.
For all these models, q and s are similar to that of the standard model, so the companion
is a cold low mass planet. For all 4 degenerate solutions, the model parameters of greatest
interest are all very similar, except for the microlensing parallax. The ecliptic degeneracy
yields nearly identical physical parameters, except that the direction of the lens-source rela-
tive motion is different. Potentially, this angle can be measured with follow-up observations
(Bennett et al. 2015; Batista et al. 2015). In contrast to the ecliptic degeneracy, the wide de-
-- 12 --
generacy implies different amplitudes, πE, of the microlensing parallax vector, which implies
different lens system masses, as discussed in Section 5 below.
Thus, this wide degeneracy presents us with two different classes of physical models,
P±∓ and P±±, where the P±± models have larger πE implying smaller lens system masses
and distance (see Section 5 and Table 8).
MOA, ∆χ2
These four parallax models are preferred over the standard πE = 0 model in both
the MOA-Red and OGLE-I bands by (∆χ2
OGLE) = (−21.8,−11.5), (−21.6,−11.1),
(−21.9,−12.5) and (−23.6,−12.8) for theP+−, P−+, P++ and P−− models, respectively. In
total, the χ2 differences range from ∆χ2 = −33.3 to −37.1. In Figure 5, one can also see that
these χ2 improvements came from the same region of the light curves around the planetary
anomaly in both datasets. Microlensing parallax signals can sometimes be mimicked by the
systematic errors in the light curve photometry, but a consistent signal seen in both the
MOA and OGLE data implies that the signal is likely to be real.
4.3. Xallarap model
The xallarap effect is a light curve distortion caused by the orbital motion of the source
star (Griest & Hu 1992; Han & Gould 1997), so it only occurs if the source star has a bi-
nary companion (Derue et al. 1999; Alcock et al. 2001). Xallarap can be represented by five
additional model parameters. The xallarap vector ξE = (ξE,N, ξE,E) is similar to the parallax
vector, πE, and represents the direction of the lens-source relative motion. The amplitude of
the xallarap vector, ξE = as/rE is the semimajor axis of the source's orbit, as, in units of the
Einstein radius projected on the source plane, rE = θEDs. The other xallarap parameters
are the direction of the observer relative to the source orbital axis, with vector components
R.A.ξ and decl.ξ, and the source binary orbital period, Tξ. For an elliptical orbit, two addi-
tional parameters are required, the orbital eccentricity, ǫ and time of perihelion, tperi, which
we did not consider here as their inclusion did not improve the fit of the model to the data.
We found xallarap models giving only marginally better χ2 values compared to parallax
models for Tξ ≥ 160 days and worse values of χ2 for shorter values of Tξ. This is not surprising
as it is known that xallarap effects can mimic parallax effects (Smith, Mao & Paczy´nski
2003; Dong et al. 2009a). Including xallarap yields a slight improvement of ∆χ2 ∼ −5 for
160 ≤ Tξ < 200 days and ∆χ2 ∼ −9 at Tξ ≥ 200 days. However, these models lead to
a xallarap amplitude of ξE ≥ 0.26, which is larger than would be induced by a "normal"
-- 13 --
main-sequence companion. Here ξE is expressed, making use of Kepler's third law, by
2
3
.
(1)
ξE =
as
rE
=
1AU
rE (cid:18) Mc
M⊙(cid:19)(cid:18) M⊙
Mc + Ms
Tξ
1yr(cid:19)
These models require a source companion of mass Mc > 6M⊙ for Tξ ≥ 160 days and
Mc > 40M⊙ for Tξ ≥ 200 days. Such a heavy object would most likely be a stellar remnant
or a black hole -- in either case, a rare object and thus an unlikely source companion. For
this reason we reject the inclusion of the xallarap in our models.
4.4. Orbital motion model
4.4.1. Linear Orbital Motion
The orbital motion of the planet around the host star causes a similar effect as parallax.
To a first-order approximation, the orbital motion of the planet is described by two parame-
ters, the rate of change, ω = dα/dt (radian yr−1), of the binary axis angle α, and the rate of
change ds/dt (yr−1), of the projected lens star and planet separation s (Dong et al. 2009a;
Batista et al. 2011), as follows,
s = s0 + ds/dt(t − tfix), α = α0 + ω(t − tfix),
(2)
where, s0 and α0 are instantaneous value of s and α at the time tfix. We required the planet
to be bound. That is, the ratio of the projected kinetic energy and potential energy,
(cid:18)KE
PE(cid:19)⊥
=
(r⊥/AU)3
8π2(M/M⊙)"(cid:18)1
s
ds
dt(cid:19)2
+(cid:18)dα
dt(cid:19)2# yr2,
(3)
which is less than the ratio of kinetic to potential energy (KE/PE) in three dimensions,
was required to be less than unity in the MCMC calculations used to determine the model
parameter distributions. The four best linear orbital motion models (with scripts "L") that
correspond to each of four parallax models in Table 2, are shown in Table 3. One finds that
πE and its uncertainty significantly increased, while the χ2 only slightly improved. This is
because of the well known degeneracy between one component of the parallax vector, πE,⊥,
which is the perpendicular to the binary axis and close to πE,N in this case, and the lens
orbital rotation on the sky, ω. As an example, ∆χ2 distribution of πE,N and ω for the model
P+−L is shown in Figure 7.
Note that there are two additional degenerate models P±±L′ which have smaller s0 ∼
1.97 and larger ds/dt ∼ −4.1 yr−1 compared to the other models. Here, s of these models
-- 14 --
are similar to others, s ∼ 2.4, when the source crosses the planetary caustic. However these
models are disfavored with the full Keplerian orbit in the following analysis.
The physical parameters of the lens system of these models are shown in Table 9 (see
details in section §5). The host stars in these four models have a brown dwarf mass. Note
that (KE/PE)⊥ of these models given in Table 9 are close to unity. The probability of having
such high value is quite low as it requires very large eccentricity of e ∼ 1, seeing the orbital
plane face-on.
If the parameters are not well constrained by the light curve, the density
distribution of the MCMC chain depends on the prior probability of the fitting parameters
in MCMC. Although the linear approximation of the lens motion is good enough in most of
the cases, this parameterization inadvertently assumed the uniform prior on all microlensing
fitting parameters, which is not physically justified. We need to use a full keplerian orbit
parameterization to introduce physically justified priors.
4.4.2. Full Keplerian Orbit
To take the proper weighting on the orbital parameters, we adopt the full Keplerian
parameterization by Skowron et al. (2011). The advantage of the full Keplerian orbit is not
only being more accurate and allowing only bound orbital solutions, it also enables us to
introduce physically justified priors on the orbital parameters. In addition to the parameters
defined above, we introduce the position and velocity along the line of sight, sz in units of
rE and dsz/dt in yr−1. Then, the three dimensional position and velocity of the secondary
relative to its host can be described by (s0, 0, sz) and s0(γk, γ⊥, γz) = (ds/dt, s0ω, dsz/dt).
We run MCMC fitting using the microlensing parameters with these six instantaneous
Cartesian phase-space coordinates, in which we transform the "microlensing" parameters to
"Keplerian" parameters, i.e., eccentricity (e), time of periapsis (tperi), semi-major axis (a)
and three Euler angles, longitude of the ascending node (Ωnode), inclination (i), and argument
of periapsis (ωperi). By following Skowron et al.
(2011), we assume flat priors on values of
eccentricity, time of periapsis, log(a), and ωperi. Owing to the fact that orbital orientation is
random in space, we multiply the prior by sin i. We must multiply the Jacobian of the pa-
rameter transformation function, jkep = ∂(e, a, tperi, Ωnode, i, ωperi)/∂(s0, α0, sz, γk, γ⊥, γz)
(Eq. B6 in Skowron et al. 2011). So we adopt the Keplerian orbit prior of Pkep = jkep sin ia−1
and added the ∆χ2 penalty of ∆χ2
kep = −2 ln(Pkep).
We first show the results with full Keplerian orbit (with scripts "K") without any
priors in Table 4 and Table 10. The results are almost same as the ones with the linear
approximation of the orbit. The large eccentricity of e ∼1 seeing the orbital plane face-on
-- 15 --
(i ∼ 0, 180◦ ) is as expected from the large (KE/PE)⊥ in the linear orbit. The physical
parameter of the keplerian orbits, semi-major axis akep, period P , e, and i are not well
constrained so that they have very large asymmetric error bars in MCMC in Table 10. Here,
when the best fit is larger or smaller than the 68% confidence interval of MCMC chains,
the upper or lower limit is designated as "+0.0" or "−0.0", respectively. So the light curve
shape itself does not constrain the parameters more than the linear orbit model, except that
it ruled out the models with smaller s0 and larger ds/dt corresponding to P±±L′. The best
fit parallax vectors are larger than that of the static model as shown in Figure 6. Note that,
the ratio of 3D kinetic to potential energy (KE/PE) can be calculated in these full Keplerian
orbit models as shown in Table 10. which are also close to unity.
The results with the Keplerian orbit with the Keplerian prior (with scripts "Kp") are
shown in Table 5, Table 11 and Figure 6. With the Keplerian prior, the circular orbits with
e ∼ 0 are preferred contrary to the large eccentricity without the Keplerian prior. This
is because Jacobian jkep is proportional to 1/(e sin i) as noted by Skowron et al.
(2011)
and thus smaller values of eccentricity are preferred. Here, technically, the lower limit of
eccentricity is set to be 10−4 to avoid a numerical problem in MCMC as suggested by
Skowron et al.
(2011).
πE is reduced by a factor of 1/2∼2/3 because the circular orbit is preferred by the
Keplerian prior. So the lens masses increased, while the hosts are still the high-mass and
low-mass brown dwarfs. As for the models P±∓Kp, there are other minima with a lower
parallax value of πE ∼ 0.2 with similar final χ2 whose host is a low-mass M-dwarf. This is
l . But χ2 values from the light curves alone
because Pkep prefers larger values of Dl by D6
are larger than brown dwarf models. So there seems to be some conflict between light curve
and prior.
4.4.3. Stellar Kinematic constraint
Here, we applied the prior for the Galactic kinematics by following Batista et al. (2011).
In Table 11, the projected lens-source relative velocity vt = (vt,l, vt,b) of these Kp models in
the Galactic coordinate differs significantly. Those of the M-dwarf models are significantly
different from the expected value from the Galactic kinematics as shown in Fig. 8. Here we
assume a source distance of Ds = 8kpc (Reid 1993; Honma et al. 2012), the proper motion
of the Galactic center is µGC = 6.1 mas yr−1 (Backer & Sramek 1999; Reid & Brunthaler
2004; Honma et al. 2012) , the proper motion dispersion of stars in the bulge is σµ,GB = 3
mas yr−1 (Kuijken & Rich 2002), the velocity dispersions of the Galactic disk stars in the
Galactic coordinates are σDisk,l = 34 km s−1 and σDisk,b = 18 km s−1 (Binney & Merrifield
-- 16 --
1998; Minchev, Chiappini & Martig 2013; Sharma et al. 2014). Then the expected average
(vt,exp,l, vt,exp,b) and dispersion (σt,l, σt,b) of the lens projected velocity are calculated. The
probability of having observed vt can be given by
Pkin = exp"−
(vt,l − vt,exp,l)2
2σ2
t,l
# exp"−
(vt,b − vt,exp,b)2
2σ2
t,b
#.
(4)
The ∆χ2 penalty of ∆χ2
kin = −2 ln(Pkin) is about +16 and +15 for M-dwarf P+−Kp and
P−+Kp models respectively. On the other hand the penalty is +9, +4, +2 and +2 for brown
dwarf P+−Kp, P−+Kp , P++Kp and P−−Kp models, respectively. So the M-dwarf models are
less preferred.
Thus, we conducted MCMC runs by adding the penalty ∆χ2
kin. The results (with scripts
"Kpk") are shown in Table 6, Table 12 and Figure 6. The model light curve of P++Kpk is
shown in Figure 2. As expected, πE values for the M-dwarf models increase to ∼ 0.3 to reduce
the ∆χ2
kin and the total χ2 value became similar or larger than the brown dwarf models. In
total, low-mass brown dwarf P±±Kpk models are slightly preferred over other models.
We adopt these Kpk models as our main results of this paper because they use the most
realistic priors and constraints. See details in §4.5.
4.4.4. Galactic mass density prior
Finally, we applied the prior for the Galactic mass density model (Batista et al. 2011;
Skowron et al. 2011).
Pgal = ν(x, y, z)f (µ)[g(M)M]
l µ4
D4
πE
,
(5)
which is the microlensing event rate multiplied by the Jacobian of the transformation from
microlensing parameters to physical coordinates, jgal = ∂(Dl, M, µ)/∂(tE, θE, πE). Here
ν(x, y, z) is the local density of lenses, g(M) is the mass function. f (µ) is the two-dimensional
probability function for a given source-lens relative proper motion, µ = vt/Dl, which is set
to unity because it is already implemented in Pkin above. We adopt the Galactic model by
Han & Gould (1995) for ν(x, y, z) and adopt g(M) ∝ M −1 by following Batista et al. (2011).
gal = −2 ln(Pgal) are shown (with
scripts "Kpkg") in Table 7, Table 13 and Figure 6. The light curves, caustics, critical curves
and source trajectory of the models comprising both parallax and planetary orbital motion
with various different priors are almost same as that of the parallax-only models as shown
The results of MCMC runs by adding a penalty of ∆χ2
-- 17 --
in Figure 2 and Figure 3. Overall, the lens masses slightly increased relative to that of Kpk
models because this Pgal prefers larger Dl, i.e., smaller πE.
In addition to above six models, there are two more minima for P±±Kpkg with a lower
parallax value of πE ∼ 0.8 with similar final χ2 values. This is also because the prior Pgal
prefers larger Dl and smaller πE values. These solutions happen to have similar πE values
with that of high-mass brown dwarf P±∓Kpkg models, hence the similar physical parameters.
As for the P±∓Kpkg models, there are two other minima which each have a much smaller
parallax value of πE ∼ 0.035 and a smaller final χ2 =15107. This is because their large
values of Dl = 7 kpc are preferred by Pgal. These solutions have a very heavy host mass
In addition, these solutions have a value of
of Mh ∼ 1.7M⊙ which would be quite rare.
χ2
lc ∼ 15224 which is larger than any other model with parallax, which conflict to the
preference by Pgal. Furthermore, these models have very bright H-band source magnitudes
Hs = 15.816± 0.017 and Hs = 15.824± 0.018 for P++Kpkg and P−−Kpkg, respectively. These
are too bright compared to the Keck AO measurement of the target, H = 15.90 ± 0.02, by
3σ (see Section 6). If we assume that the host is a main sequence star, then total brightness
of the source plus lens is expected to be brighter and ruled out by Keck measurements by
more than 4σ. For these reasons we do not consider these solutions to be real, and are not
listed amongst the other solutions in the Tables.
4.5. Model Selection
We have presented a large number of models with various high order effects and priors.
Here we summarize which of these models are preferred over the others. As discussed in §4.2,
the parallax signal looks qualitatively real because the signal come from the theoretically
expected part of the light curve and it is consistent in both MOA and OGLE-I datasets. The
χ2 improvement by the parallax-only models over the standard model are ∆χ2 = −33.3 ∼
−37.1 with 2 additional parameters, which is equivalent to a confidence level of 5.4∼5.8σ
and is formally significant.
Furthermore, we compared the models by using the common statistical criteria, Akaike's
Information Criterion (AIC), AIC = χ2 + 2nparam, and the Bayesian information criterion
(BIC), BIC = χ2 + nparam ln(Ndata) (Burnham & Anderson 2002), which includes a penalty
discouraging an overfitting. The smaller the AIC and/or BIC values are the better model
is. Here we adopt the number of data points Ndata = 2913 during the event at 6450 <
HJD′ < 6620 for BIC because the baseline data outside of this range do not constrain the
parallax signal.
-- 18 --
The differences in these criteria between the parallax-only models and standard model
are ∆AIC = −29 ∼ −33 and ∆BIC = −17 ∼ −21. Thus the parallax-only models are
better than the standard model. The parallax models with linear orbit (L) are also better
than standard model by ∆AIC = −28 ∼ −41 and ∆BIC = −5 ∼ −17. The parallax
models with full keplerian orbit with two more parameters are better than the standard in
AIC but not in BIC.
The ∆χ2 of the parallax models with linear orbit (L) relative to the parallax-only models
are only marginal of −3 ∼ −12 with two additional parameters. The differences in these
criteria are ∆AIC = +0.7 ∼ −8 and ∆BIC = +13 ∼ +4, and it is worse for the full
keplerian orbit with two more parameters. Thus, the inclusion of the orbital motion is not
justified by these criteria.
However, the reason that we must introduce the orbital motion is not to improve the
goodness of the fit, but to avoid the bias of the value and the underestimate of the uncertainty
in parallax parameters due to the known degeneracy between the parallax and the lens orbital
motion as shown in section §4.4.1.
Furthermore, even though the full Keplerian orbit does not improve the goodness of the
fit with two additional parametes, its incorporation has the following benefits. As Skowron
et al (2011) noted, in addition to being more accurate, the advantage of the full Keplerian
orbit is to avoid all unbound orbital solutions (with eccentricity > 1) and to enable the
introduction of priors on the values of orbital parameters directly into MCMC calculations.
If the uncertainty of the parameters are relatively large like in this event, the density dis-
tribution of the MCMC chain depends on the prior probability of the fitting parameters
in MCMC. However, the linear orbital motion parameterization inadvertently assumed the
uniform prior on all microlensing fitting parameters, which is not physically justified. On
the other hand, the full keplerian orbit parameterization enable us to properly weigh the
MCMC chains with physically justified priors.
We conducted the modeling with three different sets of relatively realistic priors includ-
ing the Keplerian prior, i.e., Kp, Kpk, and Kpkg. We think that the model Kpk is more realistic
than Kp as the galactic kinematics constraint is applied. The models with the Galactic mass
density prior Kpkg may also be useful, but we do not know if the assumption that the distri-
bution of the planetary systems is uniform throughout the Galaxy is valid. Thus we adopt
the Kpk models as our main results of this paper.
It is important to note that the results of all these models Kp, Kpk, and Kpkg, are
basically same within their errors, thus our main conclusion does not depend on the choice
of these priors.
-- 19 --
Among the models in Kpk, the low-mass brown dwarf models, i.e., P++Kpk and P−−Kpk,
lc, χ2 and Hs (see §6). But we accept all the modes equally
are slightly preferred by both χ2
as possible solutions.
5. Lens properties
The lens physical parameters can be derived for this event because we could measure
both the parallax and finite source effects in the light curve.
The OGLE-IV calibrated color magnitude diagram (CMD) in a 2′ × 2′ region around
the event is shown in Figure 9. Figure 9 also shows the center of the Red Clump giants
(RCGs) (V − I, I)RC,obs = (2.047, 15.73) ± (0.002, 0.04) and the model independent OGLE
V − I source color found by linear regression and the best fit source I magnitude of the
model P++Kpk, (V − I, I)s = (1.985, 18.13) ± (0.008, 0.02). Is for other models are almost
same, as shown in Table 2-7.
Assuming the source suffers the same dust extinction and reddening as the RCGs and
using the expected extinction-free RCG centroid (V − I, I)RC,0 = (1.06, 14.39) ± (0.06, 0.04)
at this position (Bensby et al. 2013; Nataf et al. 2013), we estimated the extinction-free
color and magnitude of the source as (V − I, I)s,0 = (1.00, 16.80) ± (0.06, 0.06). This color
measurement is consistent with the independent measurement of (V − I)s,0 = 1.02± 0.06 by
the CTIO telescope (see §3). We use the average of OGLE and CTIO colors, (V − I)s,0 =
1.01 ± 0.06 in the following analyses. Here the errors in (V − I)s,0 are dominated by the
error in (V − I)RC,0. These values are consistent with the source being a K2 subgiant
(Bessell & Brett 1988).
Following Fukui et al. (2015), we estimated the source angular radius, θ∗, by using the
relation between the limb-darkened stellar angular diameter, θLD, (V − I) and I given by
Equation (4) of Fukui et al. (2015). This relation is derived from a subset of the interfer-
ometrically measured stellar radii in Boyajian et al. (2014), in which the dispersion of the
relation is ∼2% by using only stars with 3900K < Teff < 7000 K to improve the fit for FGK
stars. This yield the source radius of θ∗ = θLD/2 = 1.84 ± 0.12 µas.
The spectrum of the source was taken by the UVES spectrograph on the Very Large
Telescope (VLT) at a time when the source was still magnified as a Target of Opportunity
(ToO) observation. Reductions were carried out with the UVES pipeline (Ballester et al.
2000). The observation and data analysis have been done by the same manor as Bensby et al.
(2011, 2013). This gives the source effective temperature, Teff = 4854 ± 66 K, the gravity,
in prepara-
log g = 3.30 ± 0.14, and the metallicity, [Fe/H]= −0.17 ± 0.09 (Bensby et al.
-- 20 --
tion). By using these values and the the relation by Casagrande et al. (2010), we derive the
extinction-free source colors, (V −I)s,0,spec = 1.036±0.047 and (V −H)s,0,spec = 2.244±0.078.
Thus we get (I − H)s,0,spec = 1.208 ± 0.091. The extinction-free H-band source magnitude
is given as Hs,0 = Is,0 − (I − H)s,0,spec = 15.59 ± 0.11.
The expected extinction-free (I − H)s,0 from the measured (V − I)s,0 are (I − H)s,0 =
1.188 ± 0.082 and (I − H)s,0 = 1.119 ± 0.074 by using the stellar color-color relation of
Bessell & Brett (1988) and Kenyon & Hartmann (1995), respectively. These are roughly
consistent to (I − H)s,0,spec. By a linear regression of OGLE-I-band light curve and CTIO
H-band light curve (see §3), which are calibrated to the 2MASS scale, we got the source
(I − H) color as,
(I − H)s = 2.256 ± 0.016.
(6)
Correcting the extinction by using the measured extinction and reddening (E(V − I), AI) by
RCGs and the extinction law of Chen et al. (2013), we got (I − H)s,0,OGLE,CITO = 1.259 ±
0.071, which is also consistent with (V − H)s,0,spec.
Then we got θ∗ = 1.90±0.11 µas by using the relation between θLD, Hs,0, (V −H)s,0 and
[Fe/H], given by Equation (9) of Fukui et al. (2015), which is also driven in the same way as
Equation (4) of Fukui et al. (2015) but with the metallicity term. Note H in the relation is
in Johnson magnitude system. Thus the observed H-band source magnitude which is in the
2MASS system, is converted to the Johnson system by following Fukui et al. (2015). This is
consistent with above value. The average of above values are,
θ∗ = 1.87 ± 0.12 µas,
(7)
where we adopt the larger error from the estimate with (V − I, I), conservatively. This value
is about the median of those from other models and differences from them are less than 2%,
thus we adopt this value for all models in the following analysis.
We also tested the traditional method as follows. Following Yoo et al. (2004), the dered-
dened source color and brightness (V -K, K)s,0 = (2.2, 15.6) are estimated using the observed
(V -I, I)s,0 and the color-color relation of Kenyon & Hartmann (1995). By using this (V -K,
K)s,0 relation and the empirical color/brightness-radius relation of Kervella et al. (2004), we
estimated the source angular radius, θ∗ = 1.85 ± 0.16 µas, where the error includes uncer-
tainties in the color conversion and the color/brightness-radius relations. This is consistent
with the above value.
The physical parameters of all models are listed in Table 8-13. The physical properties
of three models with realistic priors and constraints, i.e., Kp, Kpk, and Kpkg, are basically
same within the error bars. In the following analysis, we focus on the model Kpk.
-- 21 --
Here we summarize the physical properties of the lens system by showing the average
values of various models in Table 12 for clarity. The averages are taken without any weighting
by their error bars. The uncertainties are given by the maximum and minimum values of
1-σ upper and lower limits of all (or a group of) models. The angular Einstein radii, and
geocentric lens-source relative proper motion µgeo, which are independent of the parallax
values, are estimated, respectively, as follows,
θE =
µgeo =
θ∗
ρ
θE
tE
= 0.48 ± 0.06 mas,
= 8.4 ± 1.2 mas yr−1.
(8)
(9)
This µgeo is consistent with the typical value for disk lenses of µ ∼5-10 mas yr−1 (Han & Gould
1995).
The total mass and distance of the lens system can be given by M = θE/(κπE) and
Dl = AU/(πEθE + πs), where κ = 4G/(c2AU)=8.144 mas M −1
⊙ , πs =AU/Ds and Ds ∼ 8
kpc is the distance to the source (Dong et al. 2009b). Thus these quantities depend on the
parallax parameter and we have three groups of solutions in models, i.e, small πE ∼ 0.3
(P±∓Kpk), medium πE ∼ 0.8 (P±∓Kpk and P±±Kpk) and large πE ∼ 2 (P±±Kpk). The
distance to the system, Dl, the mass of the host, Mh, and planet, Mp, and their projected
separation, a⊥, of these solutions are,
Dl = 3.6+0.6
−0.8 kpc, 1.8+0.4
−0.2 kpc, or 0.85+0.13
−0.08 kpc,
Mh =
M
1 + q
= 0.19+0.05
−0.06 M⊙, 0.068+0.019
−0.011 M⊙, or 0.025+0.005
−0.004 M⊙,
Mp =
qM
1 + q
= 21+6
−7M⊕, 7.9+1.8
−1.2M⊕, or 3.2+0.5
−0.3M⊕,
a⊥ = sθEDl = 4.2+0.7
−0.9 AU, 2.1+0.4
−0.2 AU, or 0.94+0.12
−0.09 AU,
(10)
(11)
(12)
(13)
respectively. Here a⊥ is the 2-dimensional (2D) projection of a 3D elliptical orbit having a
semi-major axis a. The expected 3D semi-major axis can be estimated by aexp = p3/2a⊥
(Gould et al. 2014). The best fit 3D semi-major axis by the Keplerian orbit, akep, are in-
between of these values in the case of this event.
-- 22 --
The semi-major axis akep normalized by the snow-line, asnow = 2.7(Mh/M⊙), are
akep
asnow
= 8.9+10.5
−1.4 , 12+7
−1, or 14+11
−1 ,
(14)
The effective temperature of the planet at the time of its formation based on the host mass
and host-planet separation are also given in the tables.
The small parallax models suggest that the planet has a mass similar to Neptune (17M⊕)
orbiting a very low mass M-dwarf in the Galactic disk. The planet is very cold as the
estimated separation is 8.9+10.5
times larger than the snow-line. This is comparable to
−1.4
Neptune's semi-major axis, i.e., 11 times larger than the Sun's snow-line. This interpretation
of the planetary signal for MOA-2013-BLG-605Lb, therefore, suggests the planet is a Neptune
analog.
The medium parallax models correspond to a miniature Neptune (or large-mass "super-
Earth") orbiting a high-mass brown dwarf host. The planet is even colder as the planetary
orbit radius is 12+7
−2 times larger than the snow line, which is also similar to the Neptune.
The large parallax models correspond to three times the Earth-mass planet orbiting a
low-mass brown dwarf host. The planet is colder because the planetary orbit radius is 14+11
−1
times larger than the snow line.
These solutions of Kpk are compared to the planets found by other methods in Figure
1. As one can see in the right-hand panel of Figure 1, in either group of models, this planet
is the coldest low mass planet ever found and it is very similar to Neptune.
6. Keck AO Observations and Lens Mass constraint
We observed with the NIRC2 instrument mounted on KECK-II the microlensing target
MOA-2013-BLG-0605 on July 26, 2015. We used the Wide camera giving a pixel scale of
0.04 arcsec and a field of view of 40 arcsec. We adopted a 5 position dithering pattern, and
did 30 exposures of 10 seconds each. We performed dark subtraction and flatfielding in the
standard manner for an IR detector. We then stacked the frames using Swarp (Bertin et al.
2002), without subtracting the background. The final image is shown in Figure 10.
For absolute calibration, we used images from the VVV survey done with the VISTA
4m telescope at Paranal (Minniti et al. 2010). We extracted a 3 arcmin JHK band images
centred on the target. We computed a PSF model using PSFEX software (Bertin & Arnouts
1996), and measured fluxes on the frames using SExtractor with this PSF model. We cross
identified the stars from the field with 2MASS catalogues. We selected 300 stars that are
-- 23 --
bright while not saturated on VVV, and derive the photometric zero points with an accuracy
of 0.004 mag. We then use the VVV catalogue to perform the astrometric calibration of the
KECK frame.
We then measure the fluxes using SExtractor as described in (Batista et al. 2014). We
cross identified 39 stars in both the VVV and KECK image. We exclude the stars saturated
on the KECK, and derive the zero point of KECK photometry. In the non-AO PSF, there
are two stars, the source with H = 15.90 ± 0.02 and a blend at ∼ 0.3 arcsec to the south
with H = 17.01 ± 0.03. Here, these two stars are blended in the OGLE reference image
and the cataloged centroid is in-between of them. The actual source position during the
magnification on the OGLE difference image was precisely measured as shown by the red
cross in Figure 10. This clearly resolved the source and showed that the blend measured in
the fitting process is not the lens.
By using the apparent source color (I − H)s given by Eq. (6) and the best fit Is, H-
band source magnitude, Hs are calculated as shown in Table 8-13. There is a trend that
the smaller the parallax is, the brighter the source is. The Hs of low-mass brown dwarf
models are almost same as the Keck measurement of H = 15.90 ± 0.02 or only slightly
brighter within 1σ. The Hs of high-mass brown dwarf and most of M-dwarf models are
within 2σ. The M-dwarf models, P−+Kp, P−+Kpk and P−+Kpkg have Hs = 15.856 ± 0.022,
15.856 ± 0.024, and 15.845 ± 0.026, which are 1.6, 1.7 and 2.1 σ brighter than the Keck
measurement. Furthermore, these are 2.0, 2.0 and 2.4 σ brighter when those include the lens
(host) brightness (Kroupa & Tout 1997) of Hh = 20.96± 0.24 (Mh = 0.28M⊙, Dl = 4.3kpc),
21.22 ± 0.24 (Mh = 0.20M⊙, Dl = 3.5kpc) and 21.17 ± 0.24 (Mh = 0.21M⊙, Dl = 3.7kpc),
respectively. These comparisons of the Keck result and source magnitudes from the light
curve indicate that the lens is very faint and not detected. This is consistent with the physical
solutions from the parallax measurements that the host is a low-mass M-dwarf or a brown
dwarf mentioned in Section 5. The low-mass brown dwarf models are slightly preferred.
The Hs of the standard model is consistent to the Keck results, but we concluded that the
parallax models are better as discussed in §4.5.
As mentioned in Section 4.4.4, there are two minima with a much smaller parallax value
of πE ∼ 0.035 for the P±∓Kpkg models. In addition to the rarity of their heavy host mass
of ∼ 1.7M⊙ which might be a stellar remnant, their source magnitudes Hs = 15.824 ± 0.018
and Hs = 15.816 ± 0.017 are 2.9σ and 3.2σ brighter than the Keck measurement. So these
models are not likely real. If their host is a main sequence star, then total brightness of the
source plus lens are ruled out by Keck measurement by more than 4σ.
-- 24 --
7. Future Mass Measurement
Let us consider the prospects for resolving the degeneracy and characterizing the host
and the planet. In the first epoch of Keck AO observations, we could not detect any excess
light, which confirmed the lens is faint. If the second epoch is taken by HST or AO ob-
servations, then we may directly detect the host (or possibly its companion). We can then
measure the lens mass and distance or place a stronger upper limit on the lens mass.
In Table 8-13, the geocentric proper motions are reported as µgeo = θE/tE = 8 ∼ 9
mas yr−1. The heliocentric proper motion is given by (Janczak et al. 2010),
µhel = µgeo + v⊕,⊥
πrel
AU
,
(15)
where v⊕,⊥ = (v⊕,⊥,N , v⊕,⊥,E) = (−2.96,−8.24) km s−1 is the velocity of the Earth projected
on the plane of the sky at the peak of the event. The estimated µhel = (µhel,N, µhel,E) of each
model is shown in Table 8-13, and they are about 8 ∼ 9 mas yr−1. Hence it is clear that
the lens will be separately resolved by HST or AO observations in 5-10 years' time given
a diffraction limit of 50 mas. Or, if we do not see any luminous object, then the lens is a
sub-stellar object.
Not only the value but also the direction of expected relative proper motion would help
us to know if it was the lens or just an ambient star when we detect such star at 80mas from
the source 10 years later.
8. Discussion and Conclusion
There are three physical planetary solutions for the MOA-2013-BLG-605L system. One
comprises a Neptune-mass planet at a wide separation from a very low mass M-dwarf host
star, having a very similar temperature as Neptune when the planet was formed. The
second solution comprises a mini-Neptune around a high-mass brown dwarf which is even
colder than Neptune when it was formed. The third one is a super-Earth around a low-mass
brown dwarf.
These degenerate solutions may be resolved by future high resolution imaging of the
lens by the HST or ground-based telescopes using adaptive optics, after waiting a period of
time for the positions of the lens and the source to diverge. We may detect an M-dwarf lens
host star, but we do not expect to detect a brown dwarf host star by such direct imaging.
In either case, the host is one of the three least massive main sequence stars orbited
-- 25 --
by a planet for which the planet's mass was measured and for which the planet-host mass
ratio is q < 0.01. The other low host mass, low planet mass systems are MOA-2007-BLG-
192L (Mh = 0.084+0.015
−0.22 AU) (Bennett et al. 2008;
Kubas et al. 2012), MOA-2010-BLG-328L (Mh = 0.11 ± 0.01M⊙, Mp = 9.2 ± 2.2M⊕, r⊥ =
0.92 ± 0.16 AU) (Furusawa et al. 2013).
−0.012M⊙, Mp = 3.2+5.2
−1.8M⊕ r⊥ = 0.66+0.51
These planets found around very low mass (∼ 0.1M⊙) hosts have relatively small masses
themselves, ranging from super-Earth mass to Neptune mass. In contrast, a roughly equal
number of giant planets and planets with Neptune-mass or less have been found across the
whole mass range of host stars. This may imply that the formation of gas giants is more
difficult around very low mass stars compared to average K-M dwarf stars with masses of
∼ 0.5M⊙, which is the typical host star for microlensing planets. This is somewhat as
predicted by the core accretion model of planetary formation, but this work provides the
first observational evidence supporting this prediction.
This could be the first exoplanet around a brown dwarf with a mass measurement
having a planetary mass-ratio q < 0.03. There are three brown dwarf binaries where one of
the components is in the planetary mass regime (Choi et al. 2013; Han et al. 2013). However,
their mass ratios are large q ≥ 0.08, suggesting that their formation may be considered more
akin to binary formation than planetary formation.
−1.4 , 12+7
−1 or 14+11
The separation of the planet is very wide, 8.9+10.5
−1 times larger than the
snow line of ∼0.5(M/0.2M⊙) AU, ∼0.2(M/0.07M⊙) or ∼0.08(M/0.03M⊙) AU, respectively,
as seen in Figure 1. The effective temperature of the planet when it was formed, based on
the host mass and the planet-host separation, is ∼ 26K, ∼ 13K or ∼ 7K, the coldest planet
found to date apart from those planets found by the direct imaging method, which can
presently only find heavy gas-giant planets of more than a few Jupiter masses. The effective
temperature of these heavy gas-giants are a few hundreds K or higher due to their internal
heat. In either interpretation, planet MOA-2013-BLG-605Lb is orbiting around one of the
least massive objects found to date at a very wide separation. The planet is the coldest
exoplanet discovered so far. This is the first observed example of a Neptune-like exoplanet
in terms of mass and temperature or a super-Earth with Neptune-like temperature, which
are important factors in any planetary formation theory.
The probability of detecting such wide separation low mass planets is very low, even
by microlensing. The probability of a source crossing the planetary caustic is proportional
to the size of the planetary caustic, wc ∼ 4q1/2s−2 ∼ 0.01 (Han 2006), divided by half the
circle with radius of separation s, i.e., P ∼ 4/πq1/2s−3 ∼ 1 × 10−3 (s = 2.4). It is an order
of magnitude smaller than planets at s ∼ 1, where ∼ 10 planets with Neptune-mass or less
have been found by microlensing. This may imply that such low-mass planets with masses
-- 26 --
less than that of Neptune at a ≃ 10 asnow are as common as low-mass planets at a few times
of the snow line (Sumi et al. 2010; Gould et al. 2010; Cassan et al. 2012).
This conclusion may challenge the standard core accretion model and other formation
models (Ida & Lin 2004) which predict few low-mass planets with Neptune-like orbit at >10
asnow. More accurate measurements of the abundance and distribution of such low-mass
ice planets are very important in the study of the formation of Neptune and in the study
of planet formation mechanisms in general. The microlensing exoplanet search by NASA's
WFIRST satellite is expected to detect hundreds of low mass planets with Neptune-like orbit
and will constrain further planetary formation models.
TS acknowledges the financial support from the JSPS, JSPS23103002,JSPS24253004
and JSPS26247023. The MOA project is supported by the grant JSPS25103508 and 23340064.
The OGLE project has received funding from the National Science Centre, Poland, grant
MAESTRO 2014/14/A/ST9/00121 to AU. DPB acknowledges support from NSF grants
AST-1009621 and AST-1211875, as well as NASA grants NNX12AF54G and NNX13AF64G.
Work by IAB and PY was supported by the Marsden Fund of the Royal Society of New
Zealand, contract no. MAU1104. NJR is a Royal Society of New Zealand Rutherford
Discovery Fellow. AS, ML and MD acknowledge support from the Royal Society of New
Zealand. AS is a University of Auckland Doctoral Scholar. AG was supported by NSF grant
AST 1103471 and NASA grant NNX12AB99G. J.P.B., S.B., J.B.M. gratefully acknowledges
support from ESO's DGDF 2014. JPB & JB acknowledge the support of the Programme
National de Plan´etologie, CNRS, and from PERSU Sorbonne Universit´e. Work by C.H. was
supported by Creative Research Initiative Program (2009-0081561) of National Research
Foundation of Korea.
REFERENCES
Alard C., 2000, A&AS, 144, 363
Alard C., Lupton R. H., 1998, ApJ, 503, 325
Alcock, C., et al. 1995, ApJ, 454, 125
Alcock, C., Alaskan, R. A., Alves, D. R., et al. 2001, ApJ, 552, 259
An, J.H., Gould, A., 2001, ApJ, 563, L111
Backer, D. C. & Sramek, R. A., 1999, ApJ, 524, 805
-- 27 --
Ballester, P., Modigliani, A., & Boitquin, O., et al. 2000, The Messenger, 101, 31
Batista, V., et al. 2011, ApJ, 529, 102
Batista, V., et al. 2014, ApJ, 780, 54
Batista, V., Beaulieu, J.-P., Bennett, D.P. et al. 2015, ApJ, submitted
Bessell, M. S., & Brett, J. M. 1988, PASP, 100, 1134
Bennett, D.P. & Rhie, S.H. 1996, ApJ, 472, 660
Bennett, D. P., Anderson, J., Bond, I. A., Udalski, & A., Gould, A. 2006, ApJ, 647, 171
Bennett, D. P., Bhattacharya, A., Anderson, J., et al. 2015, ApJ, submitted
Bennett, D. P., Bond, I. A., Udalski, A., et al. 2008, ApJ, 684, 663
Bennett, D. P., Rhie, S. H., Nikolaev, S., et al. 2010, ApJ, 713, 837
Binney, J., & Merrifield, M. 1998, Galactic Astronomy (Princeton, NJ:Princeton Univ. Press)
Bensby, T., et al. 2011, A&A, 533, A134
Bensby, T., Yee, J. C., Feltzing, S., et al. 2013, A&A, 549, A147
Berta, Z. K. ; Irwin, J., Charbonneau, D., 2013, ApJ, 775, 91
Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
Bertin, E., Mellier, Y., Radovich, M., Missonnier, G., Didelon, P., & Morin, B. 2002, in
ASP Conf. Ser. 281, Astronomical Data Analysis Software and Systems XI, ed. D. A.
Bohlender, D. Durand, & T. H. Handley (San Francisco, CA: ASP), 228
Bond I. A. et al., 2001, MNRAS, 327, 868
Borucki, W. J., Koch, D. G., Batalha, N., et a. 2012, ApJ, 745, 120
Boyajian, T. S., van Belle, G., & von Braun, K. 2014, AJ, 147, 47
Burke, Christopher J. et a. 2014, ApJS, 210, 19
Burnham, K. P. & Anderson, D. R. 2002, Model Selection and Multimodel Inference (2nd
ed.), Springer
-- 28 --
Casagrande, L., Ram´ırez, I., Mel´endez, J., Bessell, M., & Asplund, M. , 2010, A&A, 512,
A54
Cassan, A., Kubas, D., Beaulieu, J.-P., et al. 2012, Nature, 481, 167
Clanton, C. & Gaudi, B.S., 2014, ApJ, 791, 91
Claret, A. 2000, A&A, 363, 1081
Chen, B. Q., Schultheis, M., Jiang, B. W., et al. 2013, A&A, 550, A42
Choi, J.-Y., Han, C., Udalski, A., et al. 2013, ApJ, 768, 129
Derue, F., Afonso, C., et al. 1999, A&A, 351, 87
Dong, S., Gould, A., Udalski, A., et al. 2009a, ApJ, 695, 970
Dong, S., Bond, I. A., Gould, A., et al. 2009b, ApJ, 698, 1826
Dressing, C. D. & Charbonneau, D, 2013, ApJ, 767, 95
Fernandez, J. A.; Ip, W.-H. 1984, Icarus, 58, 109
Fukui, A. et al. 2015, ApJ, 809, 74
Furusawa, K. et al. 2013, ApJ, 779, 91
Gaudi, B. S., Bennett, D. P., Udalski, A., et al. 2008, Science, 319, 927
Gould, A. 1992, ApJ, 392, 442
Gould, A. 1994, ApJ, 421, L75
Gould, A. et al. 2009, ApJ, 698, L147
Gould, A., et al. 2010, ApJ, 720, 1073
Gould, A., et al. 2014, Science, 345, 46
Gould, A. 2004, ApJ, 606, 319
Griest, K., & Hu, W. 1992, ApJ, 397, 362
Han, C. & Gould, A. 1995, ApJ, 449, 521
Han, C., 2006, ApJ, 638, 1080
-- 29 --
Han, C., & Gould, A. 1997, ApJ, 480, 196
Han, C. & Gould, A. 2003, ApJ, 592, 172
Han, C., Jung, Y. K., Udalski, A., et al. 2013, ApJ, 778, 38
Helled, R. & Bodenheimer, P. 2014, ApJ, 789, 69
Honma, M., et al. 2012, PASJ, 64, 136
Ida, S., & Lin, D.N.C. 2004, ApJ, 616, 567
Janczak, J. et al. 2010, ApJ, 711, 731
Kains N. et al. 2013, A&A, 552, 70
Kennedy, G. M., Kenyon, S. J., & Bromley, B. C. 2006, ApJ, 650, L139
Kennedy, G. M., Kenyon, S. J., 2008, ApJ, 673, 502
Kenyon, S. J. & Hartmann, L., 1995, ApJS, 101, 117
Kervella, P., & Fouqu´e, P. 2008, A&A,491, 855
Kervella, P., Th´evenin, F., Di Folco, E., & S´egransan, D. 2004, A&A, 426, 297
Kroupa, P. & Tout, C. A., 1997, MNRAS, 287, 402
Kubas, D., Beaulieu, J. P., Bennett, D. P., et al. 2012, A&A, 540, 78
Kuijken, K. & Rich, R. M. 2002, AJ, 124, 2054
Laughlin, G., Bodenheimer, P., & Adams, F.C. 2004, ApJ, 612, L73
Malhotra, R., 1993, Nature, 365, 819
Mayor, M., & Queloz, D. 1995, Nature, 378, 355
Minchev, I., Chiappini, C. & Martig, M 2013, A&A, 558, A9
Minniti, D., Lucas, P. W., Emerson, J. P., et al. 2010, New A, 15, 433
Montet, B. T., Crepp, J. R., Johnson, J. A., Howard, A. W., Marcy, G. W., 2014, ApJ, 781,
28
Morton, T. D. & Swift, J. J., 2014, ApJ, 791, 10
-- 30 --
Muraki, Y., et al. 2011, ApJ, 741, 22
Nataf, D. M. et al. 2013, ApJ, 769, 88
Calchi Novati S., et al. 2015a, ApJ, 804, 20
Calchi Novati S., et al. 2015b, ApJ, 814, 92
Park, B. -G. et al. 2004, ApJ, 609, 166
Refsdal, S. 1966, MNRAS, 134, 315
Reid, M. J. 1993, ARA&A, 31, 345
Reid, M. J. & Brunthaler, A., 2004, ApJ, 616, 872
Poleski, R. et al., 2014, ApJ, 795, 42
Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 124, 62
Quintana, E.V., Barclay, T., Raymond, S.N. et al. 2014, Science, 344, 277
Sharma, S., et al. 2014, ApJ, 793, 51
Schechter, L., Mateo, M., & Saha, A., 1993, PASP, 105, 1342S
Skowron, J., Udalski, A., Gould, A., et al. 2011, ApJ, 738, 87
Smith, M. C., Mao, S., & Paczy´nski, B. 2003, MNRAS, 339, 925
Smith, M. C., Mao, S. & Wo´zniak, P. R., 2002, MNRAS, 332, 962
Sumi, T. et al., 2003, ApJ, 591, 204
Sumi, T. et al., 2010, ApJ, 710, 1641
Sumi, T. et al., 2011, Nature, 473, 349
Tsapras T. et al. 2014, ApJ, 782, 48
Tomany, A. B. & Crotts, A. P., 1996, AJ, 112, 2872
Thommes, E. W., Duncan, M. J. & Levison, H. F., 1999, Natur, 402, 635
Udalski, A. 2003, Acta Astron., 53, 291
Udalski,A., Szyma´nski, M.K. and Szyma´nski, G. 2015, Acta Astronomica, 65, 1
-- 31 --
Udalski A. et al. 2015, ApJ, 799, 237
Verde L., et al, 2003 ApJS, 148,195
Wittenmyer, R. A., et al. 2014, ApJ, 783, 103
Wo´zniak P. R., 2000, Acta Astronomica, 50, 421
Yee, J. C., Shvartzvald, Y., Gal-Yam, A., et al. 2012, ApJ, 755, 102
Yoo, J., et al, 2004a ApJ, 603, 139
This preprint was prepared with the AAS LATEX macros v5.2.
-- 32 --
Table 1. Datasets used in the analysis
Dataset
Telescope
Filter
Ndata
k
emin
MOA-Red
OGLE-I
OGLE-V
CTIO-I
CTIO-V
CTIO-H
MOA-II 1.8 m
Warsaw 1.3 m
Warsaw 1.3 m
SMARTS-CTIO 1.3 m
SMARTS-CTIO 1.3 m
SMARTS-CTIO 1.3 m
R + I
I
V
I
V
H
9675
5514
64
15
15
149
1.092
1.387
1.571
1.0
1.0
1.0
0.0127
0.0109
0.0
0.0
0.0
0.0
Note. -- Ndata is the number of data points used in the analysis. k and
emin are the error scaling parameters (see details in the text).
-- 33 --
Fig. 1. -- : The distribution in planetary mass, Mp, versus the semi-major axis, a (left
panel) and a normalized by the snow-line (right panel) of discovered exoplanets by various
methods. Red circles indicate the microlensing planets. Microlensing planets for which mass
measurements have been made are indicated with filled circles. Microlensing planets where
the mass has been estimated by a Bayesian analysis are indicated with open circles. The
six model solutions for event MOA-2013-BLG-605Lb comprising parallax and the Keplerian
orbital motion with the Keplerian prior and the kinematic constraint (Kpk) are indicated by
purple filled circles. Black dots represent the radial velocity planets and blue filled squares
are transit planets. Cyan dots are transit planets found by Kepler. Magenta triangles
denote planets found via direct imaging. Green open squares denotes planets found via
timing measurements. Solar system planets are indicated by their initial. A green vertical
dashed line indicates the snow line. All models for MOA-2013-BLG-605Lb are very similar
to Neptune, when planet orbit radii are scaled to the snow line (right panel).
-- 34 --
Fig. 2. -- : Light curve of MOA-2013-BLG-605. Black, red and green points indicate MOA-
Red, OGLE-I and OGLE-V band data, respectively. Blue lines represent the parallax model
with Keplerian orbital motion, a Keplerian prior and the kinematic constraint P++Kpk which
is almost identical to all models with parallax and orbital motion. Middle and Bottom panels
show the detail of the planetary signal and its residual from the best model.
-- 35 --
Fig. 3. -- : Caustics (Red lines) of MOA-2013-BLG-605 of the best fit parallax models, P+−,
P−+, P++ and P−− in top-left, top-right, bottom-left and bottom-right panel, respectively.
The figure of each model with a orbital motion and various priors are similar. Insets show
a close-up view around the planetary caustic. Blue circles indicate the best fit source star
radius and position at HJD= 6534.6, just before crossing the planetary caustic. Blue lines
with arrows represent source star trajectories. The left and right black filled circles at y = 0
indicate the positions of primary and planet, respectively. The green lines show critical
curves.
-- 36 --
Fig. 4. -- : ∆χ2 distribution of the parallax parameters from the best fit parallax-only
models, without orbital motion, P++, P−+, P+− and P−−, from top to bottom. Black, red,
green, blue and orange dots indicate chains with ∆χ2 < 1, 4, 9, 16 and 25, respectively. The
black nots are larger than the others for clarity.
-- 37 --
standard− χ2
The difference in the cumulative χ2 values between the standard and the
Fig. 5. -- :
parallax models, χ2
parallax, is shown as a function of time in the bottom panel. This
is the case of the parallax model with an orbital motion, keplerian prior and kinematic prior,
P++Kpk (see section §4.4.3), but it is similar for all models with parallax. The corresponding
light curve is given in the top panel as a reference. Blue and cyan solid lines represent the
standard and P++Kpk models, while the cyan line almost overlaps the blue one. Black, red
and green points represent MOA-Red, OGLE-I and OGLE-V , respectively. The plots are
slightly shifted to match at the beginning of 2013 season for clarity, where the total ∆χ2 in
the baseline outside of the figure is almost zero. We can see that most of the parallax signals
come from around the planetary signal in both MOA-Red and OGLE-I as expected. This
consistency in two datasets support the reality of the parallax signal.
-- 38 --
Fig. 6. -- : The best fit parallax parameters (πE,E, πE,N) of the parallax-only model (open
circles in the left panel) and the parallax with the orbital motion (filled circles) with the
Keplerian orbit (K), the Keplerian prior (Kp), kinematic constraint (Kpk) and the Galactic
prior (Kpkg) from the left to right panels, respectively. The black, red, green and blue color-
code correspond the models with the geometry of P+−, P−+, P++ and P−−, respectively.
The linear orbital motion models (L) are omitted for clarity, because they are almost same
as the Keplerian orbit models (K).
-- 39 --
Fig. 7. -- : ∆χ2 distribution of the parallax parameter πE,N and orbital motion perimeter
ω from the best fit MCMC models of the P+−L, the parallax with a linear orbital motion
model. Black, red, green, blue and orange dots indicate chains with ∆χ2 < 1, 4, 9, 16 and
25, respectively. Here ω is in rad day−1 as actually used in the MCMC. The black nots are
larger than the others for clarity.
-- 40 --
Fig. 8. -- : Lens projected velocity vt = (vl, vb) in the Galactic coordinate from the best fit
with Keplerian orbit with the Keplerian prior. Top-left, top-right and bottom panels indicate
M-dwarf and brown dwarf P±∓Kp models and brown dwarf P±±Kp models, respectively.
Red circles are expected mean from the Galactic kinematics. Green error bars and circles
represent uncertainty due to the velocity dispersion of disk stars and bulge stars, respectively.
They are added in the quadrature in the total error shown in red error bars. Green bashed
line indicate the ecliptic North and East. The vt of M-dwarf P±∓Kp models are significantly
different from the expected value from the Galactic kinematics.
-- 41 --
Fig. 9. -- : OGLE-IV calibrated V and I Color magnitude diagram (CMD) within 2'×2'
around the event (green dots). A filed square indicates the center of Red Clump Giants.
The source position of P++Kpk model shown by the filled circle, which is almost identical
for all models, indicates that the source star is a K2 sub-giant.
-- 42 --
Fig. 10. -- : H-band Keck AO image within 5"×5" around the event. The red cross indicates
the actual source position measured during the magnification on the OGLE difference image.
The brighter star at the cross is the source with H = 15.90 ± 0.02 mag. The fainter star on
South is the blend with H = 17.01 ± 0.02 mag. This shows that the blend measured in the
fitting process is not the lens. The measured H-band flux of the source shows that the lens
is very faint and not detected.
-- 43 --
Table 2. Model Parameters. Standard and Parallax-only models.
Parameter
Standard
P+−
P−+
P++
P−−
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.056
0.008
20.47
0.13
7.563
0.099
2.762
0.105
2.304
0.010
3.0996
0.0004
3.37
0.10
0.000
0.000
0.000
0.000
0.000
0.000
18.167
0.011
18.508
0.015
15.911
0.019
15251.42
0.00
0.00
0.00
15251.42
15217
6573.050
0.009
19.93
0.32
7.932
0.186
3.460
0.211
2.391
0.019
3.1335
0.0077
3.88
0.15
-0.313
0.076
-0.252
0.107
0.401
0.088
18.117
0.024
18.580
0.037
15.861
0.029
15217.49
0.00
0.00
0.00
15217.49
15215
6573.050
0.009
20.04
0.29
-7.874
0.175
3.431
0.204
2.384
0.018
3.1534
0.0072
3.85
0.14
0.279
0.071
-0.261
0.100
0.382
0.082
18.125
0.023
18.568
0.034
15.869
0.028
15218.15
0.00
0.00
0.00
15218.15
15215
6573.051
0.009
20.10
0.32
7.907
0.185
3.530
0.202
2.393
0.020
2.9829
0.0084
3.91
0.15
1.114
0.086
-0.210
0.104
1.134
0.084
18.120
0.025
18.575
0.038
15.864
0.029
15216.43
0.00
0.00
0.00
15216.43
15215
6573.052
0.010
20.15
0.28
-7.890
0.164
3.611
0.215
2.395
0.017
3.3035
0.0080
3.95
0.14
-1.144
0.082
-0.249
0.104
1.170
0.088
18.123
0.021
18.571
0.032
15.867
0.027
15214.27
0.00
0.00
0.00
15214.27
15215
Note. -- HJD′ =HJD-2450000. The first subscript of model P, "+" and
"−" indicate the models with u0 > 0 and u0 < 0, respectively. The second
subscript indicates the sign of impact parameter to the secondary lens, i.e,
"++" and "−−" mean the source passes on the same side to the host and
lc is the χ2 from
planet. The 1σ error is given below each parameter. χ2
gal are χ2 penalty due to the
the light curve alone. ∆χ2
Keplerian, kinematic and the Galactic priors. Note that πE is not a fit
parameter.
kin and ∆χ2
kep, ∆χ2
-- 44 --
Table 3. Model Parameters. Linear orbit (L).
Parameter
P+−L
P−+L
P++L
P−−L
P++L′
P−−L′
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s0
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
ω (rad yr−1)
ds/dt
(yr−1)
sz
dsz/dt
(yr−1)
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.051
0.006
19.95
0.32
7.869
0.187
3.614
0.201
2.370
0.024
3.1445
0.0067
3.94
0.12
-1.528
0.194
-0.240
0.106
1.546
0.194
-1.145
0.132
-0.277
0.376
0.000
0.000
0.000
0.000
18.126
0.025
18.566
0.037
15.870
0.029
15212.02
0.00
0.00
0.00
15212.02
15213
6573.051
0.009
19.90
0.22
-7.894
0.128
3.422
0.260
2.353
0.063
3.1444
0.0076
3.85
0.19
1.420
0.226
-0.191
0.074
1.433
0.225
1.087
0.185
-0.403
0.588
0.000
0.000
0.000
0.000
18.122
0.017
18.572
0.026
15.866
0.024
15214.89
0.00
0.00
0.00
15214.89
15213
6573.055
0.010
21.33
0.31
7.418
0.154
4.183
0.355
2.302
0.093
2.9485
0.0058
4.06
0.19
3.731
0.213
0.070
0.092
3.732
0.213
2.459
0.210
-0.935
0.921
0.000
0.000
0.000
0.000
18.190
0.021
18.476
0.028
15.934
0.027
15204.17
0.00
0.00
0.00
15204.17
15213
6573.055
0.010
21.24
0.33
-7.452
0.169
4.913
0.474
2.451
0.081
3.3412
0.0144
4.41
0.23
-3.366
0.340
-0.138
0.114
3.369
0.337
-1.974
0.214
0.434
0.829
0.000
0.000
0.000
0.000
18.186
0.023
18.481
0.031
15.930
0.028
15204.36
0.00
0.00
0.00
15204.36
15213
6573.055
0.010
21.25
0.35
7.455
0.169
2.805
0.491
1.972
0.111
2.9614
0.0097
3.33
0.28
3.703
0.283
0.077
0.112
3.704
0.284
2.550
0.269
-4.104
1.049
0.000
0.000
0.000
0.000
18.185
0.024
18.482
0.031
15.929
0.029
15202.31
0.00
0.00
0.00
15202.31
15213
6573.055
0.010
21.22
0.34
-7.449
0.161
2.846
0.220
1.975
0.022
3.3222
0.0126
3.36
0.12
-3.326
0.318
-0.143
0.111
3.330
0.317
-2.117
0.227
-4.099
0.244
0.000
0.000
0.000
0.000
18.185
0.023
18.482
0.030
15.929
0.028
15202.52
0.00
0.00
0.00
15202.52
15213
Note. -- Notation is ame as Table 2. The solutions P±±L′ does not exist with the
full Keplerian orbit.
-- 45 --
Table 4. Model Parameters. Full keplerian orbit (K).
Parameter
P+−K
P−+K
P++K
P−−K
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s0
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
ω (rad yr−1)
ds/dt
(yr−1)
sz
dsz/dt
(yr−1)
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.051
0.010
19.98
0.22
7.856
0.134
3.789
0.196
2.423
0.018
3.1474
0.0077
4.03
0.11
-1.509
0.220
-0.252
0.112
1.530
0.219
-1.086
0.179
0.304
0.240
0.002
0.122
0.108
0.639
18.127
0.017
18.563
0.026
15.871
0.023
15212.20
0.00
0.00
0.00
15212.20
15211
6573.051
0.008
19.91
0.27
-7.886
0.157
3.663
0.211
2.418
0.026
3.1410
0.0075
3.98
0.13
1.413
0.187
-0.196
0.081
1.427
0.181
1.036
0.130
0.293
0.203
0.031
0.200
0.005
0.527
18.123
0.020
18.571
0.030
15.867
0.026
15214.92
0.00
0.00
0.00
15214.92
15211
6573.054
0.009
21.03
0.22
7.553
0.119
4.216
0.335
2.396
0.055
2.9553
0.0122
4.12
0.17
3.389
0.224
0.033
0.084
3.389
0.225
2.140
0.155
0.241
0.577
0.025
0.078
-0.413
0.758
18.171
0.015
18.501
0.020
15.915
0.022
15204.70
0.00
0.00
0.00
15204.70
15211
6573.055
0.008
21.14
0.18
-7.488
0.085
5.197
0.317
2.535
0.033
3.3528
0.0080
4.56
0.19
-3.351
0.197
-0.114
0.103
3.353
0.194
-1.744
0.156
1.431
0.356
0.602
0.288
-0.695
0.935
18.180
0.012
18.489
0.016
15.924
0.020
15204.43
0.00
0.00
0.00
15204.43
15211
Note. -- Notation is the same as for Table 2.
-- 46 --
Table 5. Model Parameters. Full keplerian orbit with keplerian prior (Kp).
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kp
P−+Kp
P+−Kp
P−+Kp
P++Kp
P−−Kp
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s0
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
ω (rad yr−1)
ds/dt
(yr−1)
sz
dsz/dt
(yr−1)
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.051
0.009
19.93
0.27
7.942
0.165
3.420
0.165
2.387
0.016
3.1311
0.0049
3.88
0.13
-0.005
0.050
-0.232
0.085
0.232
0.086
0.295
0.030
-0.001
0.011
0.365
0.613
0.008
0.066
18.115
0.021
18.583
0.032
15.859
0.026
15218.11
-66.30
0.00
0.00
15151.81
15211
6573.052
0.009
19.93
0.19
-7.944
0.120
3.330
0.118
2.374
0.009
3.1540
0.0048
3.81
0.11
0.030
0.040
-0.216
0.075
0.218
0.072
-0.256
0.012
-0.109
0.144
1.458
0.782
0.178
0.319
18.115
0.015
18.583
0.023
15.859
0.022
15218.29
-66.44
0.00
0.00
15151.85
15211
6573.050
0.009
20.01
0.31
7.862
0.177
3.436
0.132
2.374
0.025
3.1335
0.0068
3.84
0.10
-0.768
0.177
-0.256
0.090
0.809
0.147
-0.475
0.121
-0.122
0.222
1.031
1.006
0.281
0.516
18.126
0.023
18.565
0.035
15.870
0.028
15215.84
-61.96
0.00
0.00
15153.88
15211
6573.049
0.009
20.08
0.26
-7.816
0.150
3.401
0.160
2.375
0.017
3.1540
0.0037
3.82
0.12
0.758
0.118
-0.225
0.084
0.790
0.104
0.506
0.137
-0.001
0.049
0.027
0.110
0.067
0.629
18.132
0.019
18.556
0.029
15.876
0.025
15217.62
-62.08
0.00
0.00
15155.54
15211
6573.053
0.008
20.38
0.19
7.816
0.134
3.838
0.185
2.402
0.013
2.9711
0.0065
4.05
0.12
2.256
0.104
-0.107
0.090
2.259
0.103
1.074
0.047
0.005
0.046
0.098
0.076
-0.100
0.405
18.133
0.016
18.556
0.023
15.877
0.022
15210.27
-55.91
0.00
0.00
15154.35
15211
6573.053
0.009
20.63
0.16
-7.687
0.080
4.010
0.155
2.403
0.028
3.3154
0.0042
4.09
0.10
-2.244
0.117
-0.221
0.070
2.255
0.115
-1.031
0.090
0.072
0.295
0.256
0.484
-0.659
0.875
18.151
0.011
18.529
0.016
15.895
0.019
15207.90
-55.95
0.00
0.00
15151.95
15211
Note. -- Notation is the same as for Table 2.
-- 47 --
Table 6. Model Parameters. Kepler prior + kinematic constraint (Kpk)
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kpk
P−+Kpk
P+−Kpk
P−+Kpk
P++Kpk
P−−Kpk
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s0
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
ω (rad yr−1)
ds/dt
(yr−1)
sz
dsz/dt
(yr−1)
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.047
0.009
20.20
0.16
7.797
0.119
3.526
0.176
2.388
0.018
3.1269
0.0073
3.89
0.13
0.047
0.098
-0.315
0.057
0.319
0.062
0.302
0.036
0.047
0.076
1.197
1.126
-0.095
0.185
18.135
0.014
18.552
0.020
15.879
0.021
15218.49
-65.62
13.53
0.00
15166.40
15211
6573.054
0.009
19.83
0.23
-7.960
0.147
3.124
0.120
2.344
0.012
3.1568
0.0058
3.72
0.08
0.289
0.068
-0.118
0.103
0.312
0.072
0.040
0.030
-0.304
0.185
1.078
0.732
0.658
0.112
18.112
0.018
18.588
0.028
15.856
0.024
15220.05
-65.73
5.89
0.00
15160.20
15211
6573.050
0.009
19.96
0.10
7.890
0.072
3.557
0.131
2.394
0.011
3.1381
0.0027
3.92
0.11
-0.891
0.070
-0.259
0.059
0.928
0.067
-0.552
0.065
0.002
0.057
0.149
0.370
-0.048
0.406
18.122
0.007
18.571
0.011
15.866
0.018
15215.31
-61.27
7.81
0.00
15161.85
15211
6573.051
0.008
19.93
0.24
-7.905
0.153
3.400
0.158
2.382
0.021
3.1516
0.0059
3.85
0.13
0.793
0.074
-0.191
0.076
0.815
0.076
0.512
0.069
0.001
0.083
0.003
0.019
0.084
0.423
18.120
0.019
18.574
0.029
15.864
0.025
15217.48
-61.93
4.03
0.00
15159.58
15211
6573.052
0.009
20.41
0.32
7.804
0.187
3.812
0.189
2.396
0.015
2.9719
0.0103
4.03
0.13
2.242
0.245
-0.108
0.117
2.245
0.241
1.074
0.162
-0.001
0.035
0.023
0.212
0.192
0.995
18.135
0.025
18.552
0.036
15.879
0.030
15210.26
-55.94
1.91
0.00
15156.24
15211
6573.054
0.010
20.52
0.28
-7.749
0.154
4.054
0.279
2.407
0.014
3.3182
0.0102
4.13
0.16
-2.308
0.212
-0.190
0.108
2.316
0.213
-1.065
0.141
0.049
0.119
0.243
0.564
-0.493
0.908
18.142
0.020
18.542
0.029
15.886
0.026
15207.67
-55.80
1.88
0.00
15153.75
15211
Note. -- Notation is the same as for Table 2.
-- 48 --
Table 7. Model Parameters. Keplerian prior + kinematic+Galactic prior (Kpkg)
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kpkg
P−+Kpkg
P+−Kpkg
P−+Kpkg
P++Kpkg
P−−Kpkg
P++Kpkg
P−−Kpkg
t0 (HJD′)
tE (days)
u0 (10−2)
q (10−4)
s0
α0 (radian)
ρ (10−3)
πE,N
πE,E
πE
ω (rad yr−1)
ds/dt
(yr−1)
sz
dsz/dt
(yr−1)
Is (mag)
Ib (mag)
Hs (mag)
kep
χ2
lc
∆χ2
∆χ2
∆χ2
χ2
dof
kin
gal
6573.048
0.009
20.14
0.21
7.796
0.104
3.514
0.070
2.388
0.007
3.1253
0.0052
3.88
0.08
0.062
0.069
-0.288
0.069
0.295
0.072
0.303
0.034
0.033
0.020
0.840
0.629
-0.093
0.188
18.133
0.014
18.555
0.021
15.877
0.021
15219.03
-65.82
13.41
-38.35
15128.27
15211
6573.055
0.009
19.70
0.25
-8.042
0.162
3.137
0.082
2.349
0.016
3.1563
0.0063
3.75
0.09
0.278
0.063
-0.087
0.076
0.291
0.059
0.019
0.011
-0.299
0.075
1.086
0.907
0.648
0.238
18.101
0.020
18.605
0.032
15.845
0.026
15220.81
-65.90
5.14
-38.70
15121.36
15211
6573.051
0.009
20.01
0.29
7.860
0.172
3.436
0.146
2.373
0.016
3.1337
0.0066
3.84
0.12
-0.768
0.142
-0.256
0.090
0.810
0.138
-0.474
0.101
-0.120
0.028
1.128
1.291
0.255
0.324
18.127
0.022
18.565
0.033
15.871
0.027
15215.83
-61.96
9.06
-30.50
15132.43
15211
6573.052
0.009
19.88
0.28
-7.919
0.165
3.364
0.155
2.381
0.015
3.1528
0.0068
3.84
0.10
0.755
0.112
-0.166
0.098
0.773
0.110
0.485
0.050
-0.006
0.039
0.054
0.078
0.256
0.371
18.118
0.022
18.578
0.033
15.862
0.027
15217.93
-62.20
3.95
-30.95
15128.74
15211
6573.052
0.009
19.96
0.26
7.956
0.159
3.379
0.203
2.382
0.021
2.9868
0.0054
3.86
0.15
0.785
0.042
-0.166
0.097
0.802
0.049
-0.299
0.071
-0.082
0.053
0.199
0.103
0.972
0.203
18.113
0.021
18.586
0.032
15.857
0.026
15217.85
-62.01
3.82
-30.60
15129.06
15211
6573.051
0.009
20.09
0.22
-7.897
0.130
3.307
0.196
2.339
0.037
3.2979
0.0057
3.79
0.13
-0.836
0.114
-0.241
0.093
0.870
0.115
0.268
0.072
-0.508
0.417
1.335
0.756
0.891
0.548
18.121
0.017
18.573
0.025
15.865
0.023
15215.74
-61.58
8.35
-29.89
15132.64
15211
6573.052
0.008
20.43
0.30
7.778
0.160
3.754
0.240
2.390
0.042
2.9750
0.0062
3.99
0.15
2.200
0.110
-0.098
0.088
2.202
0.108
1.065
0.076
-0.006
0.470
0.084
0.180
0.108
0.923
18.138
0.022
18.548
0.032
15.882
0.027
15210.37
-56.04
1.93
-21.33
15134.94
15211
6573.052
0.008
20.64
0.25
-7.672
0.140
3.969
0.274
2.399
0.019
3.3119
0.0103
4.06
0.16
-2.201
0.168
-0.211
0.092
2.211
0.169
-1.025
0.067
0.087
0.069
0.374
0.496
-0.562
0.674
18.153
0.019
18.526
0.026
15.897
0.025
15208.07
-56.06
2.02
-21.20
15132.83
15211
Note. -- Notation is the same as for Table 2.
-- 49 --
Table 8. Lens physical parameters (Parallax only).
Parameter
P+−
P−+
P++
P−−
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
vt,l (km s−1)
vt,b (km s−1)
Teff,exp (K)
0.482±0.036
8.61± 0.67
-6.83± 0.52
-5.73± 0.42
3.14+0.44
−0.48
0.147+0.031
−0.037
16.99+2.99
−3.59
3.62+0.51
−0.60
9.10+1.59
−0.83
4.43+0.62
−0.73
11.14+1.95
−1.01
-208.5± 35.3
46.7± 31.7
20+5
−7
0.486±0.036
8.63± 0.67
6.19± 0.49
-6.22± 0.46
3.22+0.43
−0.49
0.156+0.030
−0.040
17.83+2.93
−3.87
3.73+0.50
−0.61
8.85+1.61
−0.76
4.57+0.61
−0.75
10.84+1.97
−0.93
71.7± 34.9
228.9± 33.8
21+5
−7
0.478±0.036
8.47± 0.66
7.98± 0.65
-2.51± 0.12
1.50+0.15
−0.10
0.052+0.008
−0.005
6.09+0.83
−0.45
1.71+0.15
−0.10
12.28+0.77
−0.78
2.10+0.18
−0.12
15.04+0.94
−0.95
63.1± 4.9
62.2± 3.0
10+1
−3
0.474±0.035
8.38± 0.64
-8.54± 0.63
-2.75± 0.14
1.47+0.14
−0.10
0.050+0.007
−0.005
5.98+0.75
−0.44
1.67+0.13
−0.11
12.44+0.83
−0.67
2.04+0.16
−0.13
15.24+1.01
−0.82
-64.7± 4.8
-9.3± 2.9
10+1
−3
Note. -- aexp = p3/2a⊥ is an expected 3D separation for the measured
projected separation, a⊥. asnow = 2.7(Mh/M⊙) is the snow-line. Teff,exp is the
effective temperature of the planet when it was formed based on the Mh and aexp.
-- 50 --
Table 9. Lens physical parameters. Linear orbit (L)
Parameter
P+−L
P−+L
P++L
P−−L
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
vt,l (km s−1)
vt,b (km s−1)
Teff,exp (K)
KE/PE⊥
0.475±0.034
8.48± 0.63
-8.83± 0.63
-2.59± 0.10
1.16+0.12
−0.17
0.038+0.004
−0.006
4.54+0.55
−0.74
1.31+0.11
−0.19
12.87+0.63
−0.46
1.60+0.13
−0.23
15.76+0.77
−0.56
-47.1± 6.1
-7.2± 3.6
9+1
−3
1.000
0.485±0.039
8.69± 0.72
8.17± 0.72
-2.37± 0.10
1.22+0.40
−0.04
0.042+0.015
−0.003
4.74+1.79
−0.21
1.39+0.43
−0.05
12.40+1.17
−0.73
1.70+0.53
−0.06
15.19+1.43
−0.90
53.5± 8.2
49.8± 4.8
9+2
−3
0.994
0.460±0.037
7.69± 0.64
6.61± 0.64
-2.84± 0.01
0.54+0.06
−0.05
0.015+0.002
−0.000
2.11+0.23
−0.23
0.58+0.05
−0.05
14.08+0.51
−1.73
0.70+0.06
−0.06
17.24+0.62
−2.12
24.4± 1.1
23.3± 0.6
5+1
−1
0.988
0.424±0.035
7.11± 0.61
-7.99± 0.61
-2.77± 0.03
0.64+0.08
−0.05
0.015+0.005
−0.001
2.53+0.36
−0.24
0.67+0.11
−0.04
16.05+0.24
−2.15
0.82+0.13
−0.05
19.66+0.30
−2.63
-16.1± 2.1
1.7± 1.2
5+1
−1
0.964
Note. -- Notation is the same as for Table 8.
-- 51 --
Table 10. Lens physical parameters. Keplerian orbit (K)
Parameter
P+−K
P−+K
P++K
P−−K
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
akep (AU)
akep/asnow
P (yr)
e
i
vt,l (km s−1)
vt,b (km s−1)
Teff,kep (K)
KE/PE⊥
KE/PE
0.464±0.032
8.28± 0.60
-8.61± 0.59
-2.59± 0.10
1.20+0.39
−0.03
0.037+0.014
−0.001
4.70+1.69
−0.16
1.35+0.40
−0.03
13.39+0.51
−0.77
1.65+0.49
−0.04
16.40+0.63
−0.94
114+0
1134+0
6304+0
−6280
0.99+0.00
−0.52
177.7+0.0
−21.9
-47.8± 7.0
-6.9± 4.2
1.0+0.5
−0.6
0.992
0.994
−1116
−112
0.469±0.034
8.40± 0.63
7.90± 0.62
-2.31± 0.09
1.26+0.27
−0.08
0.040+0.010
−0.004
4.93+1.37
−0.23
1.43+0.28
−0.08
13.10+0.92
−0.53
1.75+0.34
−0.09
16.05+1.13
−0.65
117+0
1069+0
6260+0
−6217
0.99+0.00
−0.32
0.7+22.5
−0.0
53.6± 6.7
50.1± 3.9
1.1+0.5
−0.5
0.994
0.994
−1035
−113
0.454±0.035
7.70± 0.61
6.73± 0.61
-2.60± 0.01
0.60+0.06
−0.05
0.016+0.003
−0.001
2.31+0.28
−0.16
0.65+0.04
−0.04
14.72+0.53
−1.22
0.80+0.05
−0.05
18.03+0.64
−1.50
−1006
49+0
−44
1092+0
2635+0
−2572
0.99+0.00
−0.15
4.7+7.8
−1.2
26.5± 1.4
24.7± 0.8
0.7+0.3
−0.2
0.987
0.993
0.410±0.031
6.91± 0.54
-7.76± 0.54
-2.62± 0.02
0.67+0.07
−0.05
0.015+0.002
−0.001
2.60+0.35
−0.17
0.69+0.05
−0.04
17.10+0.47
−1.29
0.85+0.06
−0.05
20.95+0.58
−1.58
−1022
44+0
−41
1080+0
2362+0
−2331
0.98+0.00
−0.25
161.6+8.5
−4.6
-16.3± 1.2
1.4± 0.7
0.7+0.3
−0.3
0.944
0.992
Note. -- Notation is the same as for Table 8, except Teff,kep is based on akep.
When the best fit is larger or smaller than the 68% confidence interval of MCMC
chains, the upper or lower limit is designated as "+0.0" or "−0.0", respectively.
-- 52 --
Table 11. Lens physical parameters. Keplerian orbit with Keplerian prior (Kp)
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kp
P−+Kp
P+−Kp
P−+Kp
P++Kp
P−−Kp
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
akep (AU)
akep/asnow
P (yr)
e
i
vt,l (km s−1)
vt,b (km s−1)
Teff,kep (K)
KE/PE⊥
KE/PE
0.482±0.035
8.62± 0.65
-0.27± 0.02
-8.81± 0.65
4.22+0.63
−0.97
0.256+0.061
−0.094
29.13+6.53
−9.98
4.86+0.75
−1.12
7.05+2.12
−0.86
5.95+0.92
−1.38
8.63+2.59
−1.05
4.92+1.72
−0.75
7.14+4.70
−0.77
21.6+11.7
−1.8
0.00+0.24
−0.00
8.7+23.7
−0.8
-188.7± 69.3
333.8±120.8
33+12
−10
0.495
0.500
0.491±0.034
8.78± 0.63
1.13± 0.09
-8.88± 0.63
4.31+0.80
−0.78
0.276+0.095
−0.083
30.68+10.12
−8.71
5.02+0.94
−0.92
6.73+1.72
−1.14
6.15+1.15
−1.13
8.24+2.10
−1.40
5.90+6.10
−1.30
7.90+9.22
−1.66
27.2+51.5
−7.1
0.00+0.59
−0.00
143.7+9.6
−13.2
-144.1± 67.2
382.3±114.3
33+10
−20
0.394
0.500
0.487±0.034
8.66± 0.63
-8.47± 0.60
-3.43± 0.20
1.93+0.40
−0.23
0.074+0.020
−0.012
8.45+2.37
−1.18
2.23+0.42
−0.26
11.17+0.99
−0.79
2.73+0.51
−0.31
13.69+1.21
−0.97
2.43+3.49
−0.19
12.18+17.62
−1.22
13.9+38.6
−1.6
0.00+0.62
−0.00
152.3+4.0
−24.4
-99.5± 16.3
-8.2± 10.6
14+2
−10
0.432
0.500
0.489±0.035
8.68± 0.64
8.08± 0.62
-3.14± 0.18
1.95+0.45
−0.11
0.076+0.018
−0.007
8.62+2.61
−0.44
2.27+0.43
−0.13
11.07+0.80
−0.59
2.78+0.53
−0.16
13.55+0.98
−0.72
2.27+0.88
−0.22
11.08+3.86
−1.86
12.4+7.1
−2.0
0.00+0.43
−0.00
3.2+49.0
−0.0
80.7± 12.2
91.8± 7.8
15+1
−4
0.499
0.500
Note. -- Notation is the same as for Table 10.
0.462±0.033
8.07± 0.59
7.41± 0.59
-2.19± 0.03
0.86+0.07
−0.06
0.025+0.003
−0.002
3.21+0.30
−0.27
0.95+0.04
−0.04
14.02+0.62
−0.53
1.16+0.05
−0.05
17.17+0.75
−0.66
0.95+0.09
−0.03
14.02+1.67
−0.50
5.8+0.8
−0.2
0.00+0.09
−0.00
3.2+9.8
−0.0
37.0± 1.5
33.1± 0.9
7+1
−1
0.498
0.500
0.458±0.031
7.90± 0.56
-8.51± 0.56
-2.57± 0.05
0.86+0.04
−0.09
0.025+0.002
−0.003
3.33+0.19
−0.40
0.95+0.01
−0.09
14.13+0.64
−0.38
1.16+0.01
−0.11
17.31+0.78
−0.47
0.96+0.72
−0.01
14.21+12.66
−0.00
5.9+8.3
−0.0
0.00+0.45
−0.00
163.9+3.8
−18.9
-28.6± 1.7
-2.4± 1.0
7+1
−3
0.464
0.500
-- 53 --
Table 12. Lens physical parameters. Keplerian prior + kinematic constraint (Kpk)
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kpk
P−+Kpk
P+−Kpk
P−+Kpk
P++Kpk
P−−Kpk
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
akep (AU)
akep/asnow
P (yr)
e
i
vt,l (km s−1)
vt,b (km s−1)
Teff,kep (K)
KE/PE⊥
KE/PE
0.481±0.035
8.48± 0.63
1.16± 0.09
-8.65± 0.63
3.59+0.60
−0.39
0.185+0.054
−0.034
21.77+5.31
−3.34
4.13+0.74
−0.48
8.25+1.26
−1.09
5.06+0.91
−0.59
10.11+1.55
−1.34
4.62+4.71
−0.71
9.23+10.23
−1.47
23.0+42.8
−4.5
0.00+0.48
−0.00
27.8+20.4
−13.1
-92.4± 26.5
263.4± 44.8
25+5
−15
0.439
0.500
0.503±0.034
9.03± 0.63
8.26± 0.59
-3.69± 0.24
3.55+0.20
−0.78
0.198+0.007
−0.065
20.58+0.56
−6.66
4.18+0.23
−0.96
7.83+1.79
−0.20
5.12+0.29
−1.17
9.59+2.20
−0.24
4.60+2.66
−1.18
8.63+7.94
−1.08
22.2+24.9
−6.1
0.00+0.51
−0.00
83.4+4.1
−4.4
177.3± 53.3
233.3± 36.5
27+9
−8
0.086
0.500
0.477±0.033
8.51± 0.61
-8.45± 0.59
-3.14± 0.17
1.76+0.19
−0.15
0.063+0.006
−0.006
7.47+0.87
−0.77
2.01+0.15
−0.14
11.82+0.70
−0.44
2.46+0.19
−0.17
14.47+0.86
−0.54
2.01+1.10
−0.04
11.83+7.08
−0.33
11.4+10.7
−0.3
0.00+0.38
−0.00
175.9+0.0
−27.0
-85.5± 5.7
-9.2± 3.6
13+1
−4
0.498
0.500
0.486±0.035
8.68± 0.65
8.19± 0.63
-2.72± 0.15
1.92+0.28
−0.12
0.073+0.014
−0.006
8.28+1.35
−0.67
2.22+0.29
−0.12
11.25+0.76
−0.72
2.72+0.35
−0.15
13.78+0.93
−0.88
2.22+0.84
−0.17
11.26+3.37
−0.93
12.2+6.7
−1.1
0.00+0.27
−0.00
3.9+34.0
−0.0
82.8± 8.6
86.1± 5.3
14+1
−4
0.498
0.500
Note. -- Notation is the same as for Table 10.
0.464±0.033
8.11± 0.61
7.44± 0.61
-2.20± 0.03
0.86+0.12
−0.06
0.025+0.004
−0.003
3.23+0.55
−0.29
0.95+0.11
−0.08
13.90+0.90
−0.44
1.17+0.14
−0.09
17.02+1.10
−0.54
0.95+0.32
−0.03
13.90+5.58
−0.78
5.8+3.1
−0.2
0.00+0.30
−0.00
4.3+28.4
−0.0
37.2± 3.6
33.2± 2.0
7+1
−2
0.497
0.500
0.452±0.034
7.86± 0.61
-8.48± 0.61
-2.47± 0.05
0.85+0.08
−0.08
0.024+0.004
−0.003
3.24+0.38
−0.30
0.93+0.08
−0.08
14.34+0.73
−0.75
1.14+0.09
−0.09
17.57+0.89
−0.91
0.93+0.68
−0.01
14.42+10.92
−0.00
5.8+7.4
−0.0
0.00+0.41
−0.00
167.7+0.0
−23.8
-27.7± 2.9
-2.6± 1.7
7+1
−3
0.480
0.500
Table 13. Lens physical parameters. Keplerian prior + kinematic + Galactic density prior (Kpkg)
Parameter
M-dwarf
high-mass brown dwarf
low-mass brown dwarf
P+−Kpkg
P−+Kpkg
P+−Kpkg
P−+Kpkg
P++Kpkg
P−−Kpkg
P++Kpkg
P−−Kpkg
θE (mas)
µgeo (mas yr−1)
µhel,N (mas yr−1)
µhel,E (mas yr−1)
Dl (kpc)
Mh (M⊙)
Mp (M⊕)
a⊥ (AU)
a⊥/asnow
aexp (AU)
aexp/asnow
akep (AU)
akep/asnow
P (yr)
e
i
vt,l (km s−1)
vt,b (km s−1)
Teff ,kep (K)
KE/PE⊥
KE/PE
0.482±0.033
8.53± 0.60
1.71± 0.13
-8.59± 0.59
3.74+0.68
−0.55
0.201+0.053
−0.045
23.54+6.29
−5.07
4.31+0.72
−0.65
7.95+1.41
−0.99
5.28+0.88
−0.79
9.73+1.72
−1.21
4.57+3.03
−0.76
8.42+5.38
−1.28
21.8+22.4
−3.6
0.00+0.37
−0.00
20.8+15.0
−6.3
-83.2± 37.0
290.9± 60.9
27+6
−11
0.464
0.500
0.498±0.034
9.01± 0.64
8.51± 0.61
-2.96± 0.19
3.70+0.49
−0.55
0.210+0.033
−0.050
21.96+3.84
−4.88
4.33+0.54
−0.66
7.64+1.47
−0.62
5.31+0.67
−0.81
9.36+1.80
−0.76
4.78+3.10
−1.75
8.43+6.10
−2.52
22.8+25.3
−10.4
0.00+0.83
−0.00
86.7+3.0
−1.2
210.0± 51.7
234.5± 33.2
28+8
−10
0.081
0.500
0.487±0.035
8.66± 0.65
-8.47± 0.61
-3.42± 0.20
1.93+0.38
−0.19
0.074+0.022
−0.010
8.45+2.26
−1.04
2.22+0.44
−0.22
11.17+0.82
−0.96
2.72+0.54
−0.28
13.68+1.01
−1.18
2.46+2.15
−0.13
12.37+9.35
−1.19
14.2+20.0
−1.2
0.00+0.40
−0.00
151.2+4.8
−25.1
-99.4± 15.3
-8.3± 10.0
14+2
−7
0.430
0.500
0.487±0.034
8.72± 0.63
8.28± 0.62
-2.53± 0.14
2.00+0.20
−0.26
0.077+0.009
−0.016
8.66+1.07
−1.46
2.31+0.19
−0.34
11.08+1.25
−0.40
2.83+0.23
−0.41
13.57+1.53
−0.48
2.31+0.38
−0.27
11.09+2.92
−0.46
12.7+2.9
−1.2
0.00+0.16
−0.00
12.6+11.4
−7.1
88.7± 13.8
88.6± 8.4
15+3
−2
0.477
0.500
0.485±0.036
8.65± 0.67
8.22± 0.66
-2.47± 0.14
1.95+0.16
−0.16
0.074+0.007
−0.009
8.35+0.75
−0.71
2.25+0.14
−0.17
11.22+0.93
−0.54
2.75+0.17
−0.21
13.74+1.14
−0.66
2.25+0.70
−0.27
11.26+3.72
−0.92
12.4+6.1
−1.7
0.00+0.31
−0.00
126.0+8.0
−15.2
86.0± 5.8
84.9± 3.5
14+2
−3
0.175
0.500
0.493±0.036
8.74± 0.66
-8.67± 0.63
-3.17± 0.18
1.81+0.38
−0.08
0.070+0.023
−0.005
7.67+2.40
−0.28
2.08+0.47
−0.08
11.09+0.71
−1.00
2.55+0.58
−0.10
13.58+0.87
−1.22
2.40+2.86
−0.45
12.76+12.25
−3.20
14.1+28.5
−4.0
0.00+0.69
−0.00
63.1+4.2
−29.4
-91.0± 11.1
-10.7± 7.0
13+1
−9
0.196
0.500
0.469±0.035
8.17± 0.63
7.52± 0.63
-2.16± 0.03
0.86+0.10
−0.04
0.026+0.005
−0.000
3.27+0.43
−0.20
0.97+0.08
−0.02
13.73+0.25
−1.26
1.19+0.10
−0.02
16.81+0.31
−1.54
0.97+0.64
−0.00
13.74+7.07
−0.42
5.9+6.1
−0.0
0.00+0.48
−0.00
3.1+28.3
−0.0
37.8± 1.7
33.5± 1.0
8+0
−3
0.499
0.500
0.461±0.035
7.96± 0.62
-8.56± 0.62
-2.53± 0.06
0.87+0.07
−0.08
0.026+0.003
−0.004
3.39+0.29
−0.35
0.97+0.04
−0.08
13.99+1.00
−0.51
1.18+0.05
−0.10
17.13+1.23
−0.62
0.98+0.29
−0.06
14.16+4.96
−0.21
6.0+2.9
−0.3
0.00+0.19
−0.00
164.3+3.5
−17.2
-29.2± 2.5
-2.8± 1.5
7+1
−1
0.470
0.500
Note. -- Notation is the same as for Table 10.
--
5
4
--
|
1507.02655 | 1 | 1507 | 2015-07-09T18:59:53 | Transiting Exoplanet Simulations with the James Webb Space Telescope | [
"astro-ph.EP"
] | In this white paper, we assess the potential for JWST to characterize the atmospheres of super-Earth exoplanets, by simulating a range of transiting spectra with different masses and temperatures. Our results are based on a JWST simulator tuned to the expected performance of the workhorse spectroscopic instrument NIRSpec, and is based on the latest exoplanet transit models by Howe & Burrows (2012). This study is especially timely since the observing modes for the science instruments on JWST are finalized (Clampin 2010) and because NASA has selected the TESS mission as an upcoming Explorer. TESS is expected to identify more than 1000 transiting exoplanet candidates, including a sample of about 100 nearby (<50 pc) super- Earths (Ricker et al. 2010). | astro-ph.EP | astro-ph | Dec 2013
Transiting Exoplanet Simulations with the James Webb Space Telescope
Natasha Batalha1, Jason Kalirai1, Jonathan Lunine2, Mark Clampin3, Don Lindler3
1Space Telescope Science Institute, 3700 San Martin Dr. Baltimore, MD 21218
2Department of Astronomy, Cornell University, 610 Space Science Building, Ithaca, NY 14853
3Goddard Space Flight Center, 8800 Greenbelt Rd, Greenbelt, MD, 20771
The James Webb Space Telescope (JWST) was ranked as the top science priority in the
Introduction
2000 Astronomy & Astrophysics Decadal Survey. The telescope has a 6.5 meter primary
aperture with sensitivity from 0.6 to 28 microns and has four science instruments with sensitivity
between 0.6 and 28 microns. Offering orders of magnitude gains in sensitivity and imaging and
spectroscopic resolution, JWST is expected to achieve breakthroughs in many astrophysical
disciplines.
Since the year 2000, our basic understanding of the demographics of planets around
nearby stars has been completely transformed. Today, new exoplanet candidates are being
discovered almost every day. The most significant result comes from the Kepler mission, which
has now accumulated over 2000 planet candidates (Batalha et al. 2013). The Kepler data
demonstrate an abundant population of “super-Earths” orbiting low mass, nearby stars. Of
particular interest are systems such as Gliese 667C (Anglada-Escudé et al. 2012), Gliese 581
(Vogt et al. 2010), Tau Ceti (Tuomi et al. 2013), HD 85512 (Pepe et al. 2011) and HD 40307
(Mayor et al. 2009), all of which are closer than 15 pc and believed to harbor sub-Neptunian
habitable zone planets.
The discovery of exoplanets around nearby stars opens the door to a new scientific front:
exoplanet characterization. Both Hubble and Spitzer have made initial contributions in this area
by observing atmospheres of planets with large hydrogen-helium envelopes, Jupiter-analogs, or
planets with hot surface temperatures (Crouzet et al. 2012, Sing et al. 2011, Deming et al. 2007).
Unfortunately, these telescopes lack the power and resolution to extend this exciting work to
super-Earth planets with M = 1-10 M⊕.
super-Earth exoplanets, by simulating a range of transiting spectra with different masses and
temperatures. Our results are based on a JWST simulator tuned to the expected performance of
the workhorse spectroscopic instrument NIRSpec, and is based on the latest exoplanet transit
models by Howe & Burrows (2012). This study is especially timely since the observing modes
for the science instruments on JWST are finalized (Clampin 2010) and because NASA has
selected the TESS mission as an upcoming Explorer. TESS is expected to identify more than
1000 transiting exoplanet candidates, including a sample of about 100 nearby (<50 pc) super-
Earths (Ricker et al. 2010).
section 2 we present the atmospheric models we used as input to the simulator. Section 3
provides the method used for determining the S/N of the simulated observations, and Section 4
summarizes the results. Section 5 provides an outline for future work that can be done with the
simulator.
In this white paper, we assess the potential for JWST to characterize the atmospheres of
In the following section 1, we provide a brief explanation of the JWST simulator. In
Dec 2013
1. JWST Simulator
We use a JWST simulator that was developed by Mark Clampin and Don Lindler at
GSFC. The JWST Simulator is comprised of several IDL routines, and takes input planetary
transit depths and converts them to an output transit spectrum as would be observed by JWST.
The simulator accepts a set of input parameters that include: the number of observed transits,
duration of each transit, jitter and drift specifications, as well as a primary in and out of transit
spectra. Line of sight requirements were under revision in the timeframe of this work and might
need to be revisited for future studies. The out-of-transit model is simply a stellar spectrum. The
in-transit spectrum is the same stellar spectrum convolved with a planetary absorption spectrum.
The simulation begins by converting the input models from wavelength vs. flux (Å vs
ergs/cm^2/sec/Å) to wavelength vs. count rates in 1/10 pixel bins in the dispersion direction
using combined efficiencies, telescope effective area and wavelength dispersion. These can then
be mapped onto an image sampled at 1/10 of a pixel by convolving with the Point Spread
Function and multiplying by the Pixel Response Function(PRF). The PRF gives insights into
how the pixels respond to starlight during a nominal observation and is based on real
measurements of the FGS detectors. Several sources of error are then added to the image,
including zodiacal and stray light, flat field errors, Poisson and read-noise. This procedure is
repeated for each observation.
The NIRSpec simulator is split up into three bands: 1 – 1.9µm, 1.5 – 3µm, 2.75 – 5µm.
The process of simulating a full spectrum requires large computing times. Each transit takes ~ 1
hour to simulate on a 3.4 GHz single processor with four cores. To extract the spectra, an
average background is computed using the region outside of the extraction region. An optimal
extraction procedure could improve the resultant spectrum, however, we do not consider optimal
extraction techniques in this work.
2. Models
spectra of super-Earths by Howe & Burrows (2012), and theoretical stellar spectra by Castelli
and Kurucz (2004). Both sets of models are publically available. Howe & Burrows generic
transit models are computed for a silicate-iron planet with the physical parameters of the GJ1214
system outlined in Table 1 and 2, except as specified. All of the simulations are computed for a
planet orbiting an M-dwarf at 0.2 M!. They compute pure atmosphere models with compositions
of water, carbon dioxide and methane as well as hydrogen rich atmosphere models for which
solar abundances are multiplied by a single factor (which can be less than or greater than unity).
To produce the primary in and out of transit models, we use new theoretical transit
Properties of Transiting Super-Earth GJ1214b (Charbonneau et al. 2009)
Radius
(R⊕)
Mass
(M⊕)
(deg)
a
e
Period
1.5803925 88.62!!.!"!!.!"
(d)
i
2.678±0.13
6.55±0.98
<0.27
Table 1
(AU)
0.0143
Table 2
Dec 2013
Properties of Host Star GJ 1214
mV
14.67
Radius
(R!)
0.2110
Mass
(M!)
0.157
Teff
(K)
3026
Luminosity
0.00328
(L!)
[Fe/H]
+0.392
We use three categories of super-Earths adapted from Miller-Ricci et al. (2009). The first
category includes super-Earths with large surface gravities that are able to retain massive
hydrogen atmospheres. We refer to this as the H-rich case. On the other end of spectrum we have
a category of super-Earths that bear close resemblance to Earth, with atmospheres depleted in
hydrogen and composed of mostly heavier molecules such as CO2 and H2O. We refer to this as
the hydrogen poor case. A third case would be the super-Earths that have lost moderate levels of
their atmospheric hydrogen due to either incomplete escape of hydrogen and/or outgassing of a
significant secondary H2 atmosphere (Miller-Ricci et al. 2009). Howe & Burrows do not have a
model that represents this intermediate case but such objects are modeled in the literature.
Therefore, the intermediate region is left for future work.
For a H-rich atmosphere, we use a 3x solar metallicity model which has a mean
molecular weight of ~2.3 g/mole and, therefore, a relatively large scale height (Earth’s MMW is
28.97 g/mole). For an atmosphere depleted in hydrogen and composed of mainly heavy
elements, we use a pure water atmosphere with a mean molecular weight of 18 and therefore, a
relatively small scale height. Refraction, which limits the altitude below which primary transits
provide spectral information (García-Muñoz et al., 2012), can be neglected here as all transiting
planets are assumed to be close to their parent stars, minimizing the effect.
3. Parameterizing the Capabilities of NIRSpec
For each atmosphere, we ran several simulations varying planet mass, planet temperature
and distance from Earth separately. These parameters are outlined in Table 3. This allowed us to
quantify the effect of each parameter on the resulting output spectrum, characterized through its
signal-to-noise ratio (SNR). Then we interpolated over the full parameter space, which enabled
us to estimate how well NIRSpec will be able to resolve whether a super-Earth has a hydrogen-
rich or a hydrogen-poor atmosphere. It also helped us determine lower limits to how small, cool,
or distant an object can be observed with NIRSpec.
Howe & Burrows Input Parameter Choices and Distances
Atmosphere
Temp (K)
Distance (pc)
Table 3
Mass (M⊕)
[Radius (R⊕)]*
400 700 1000
400 700 1000
1[1] 2[1.2] 4[1.6] 7[1.9] 10[2.2]
1[1] 2[1.2] 4[1.6] 7[1.9] 10[2.2]
6 15 30
6 15 30
*Howe & Burrows (2013) normalize their models to radius. Therefore, we assume an Earth density to
3xSolar
H2O
calculate radius (!!!).
For the H-rich atmosphere looked at a H2O/CH4 absorption features centered at 1.4 µm,
H2O at 1.8 µm and 2.7 µm, and CH4 at 4.6 µm. For the H-poor atmosphere we looked at H2O
absorption features at 1.4 µm, 1.8 µm and 2.7 µm. Since we aim to calculate the SNR for many
models, we focus initially on the first band (1 – 1.9µm) of the spectra for the H-poor model and
Dec 2013
the second band (1.5 – 3µm) of the spectra for the H-rich model. This reduces the computing
time for each simulation by one third.
To calculate the SNR of the spectra we fit the continuum using a standard first order
polynomial scheme to remove the underlying slope in the specified wavelength region. Then we
sum the flux contribution at each wavelength to get total signal and divide by the standard
deviation of the noise times the square root of the number of points (RMS noise).
R
N
S
R
N
S
25
20
15
10
0
60
50
40
30
20
R
N
S
50
40
30
20
50
Increase in SNR for H-Rich Atm
T=400K, D=6 pc
} SNR Scatter
2
4
8
Mass (Earth Masses)
6
10
M=10Me, D=6 pc
M=4Me, D=6 pc
} SNR Scatter
400
600
800
Temperature (K)
1000
T=1000k, M=4 Me
100
150
} SNR Scatter
200
250
Distance (pc)
300
R
N
S
R
N
S
R
N
S
12
10
8
6
4
2
0
16
14
12
10
8
6
16
14
12
10
8
6
4
50
Increase in SNR for H-Poor Atm
T=400K, D=6 pc
} SNR Scatter
2
4
8
Mass (Earth Masses)
6
10
M=4Me, D=6 pc
} SNR Scatter
1000
400
600
800
Temperature (K)
T=1000k, M=4 Me
100
150
} SNR Scatter
250
200
Distance (pc)
300
Figure 1. SNR vs Planet Parameters- SNR is plotted versus various planet parameters
including planet mass, planet temperature and distance from Earth. Simulations are for NIRSpec
at R~300, 25 transits, for different planets orbiting an M-dwarf. SNR calculations are described
in Section 3. Diamonds represent calculated values and the red line is a first order polynomial fit
through the points. Shown in the bottom right hand corner of each panel is the variation in SNR
across the entire wavelength region (1 – 5 µm). The middle left hand plot shows two lines
corresponding different masses to demonstrate that the slope of the temperature relationship is
not strongly affected.
The sensitivity of NIRSpec is wavelength dependent, and, therefore, the SNR differs
between the different bands defined above. We compute the SNR for each band separately. SNR
for the entire spectrum is roughly √3 times the SNR for a single band since more features are
included. Figure 1 shows the simulated spectra as a function of increasing mass, temperature and
Dec 2013
4. Simulation/SNR Results
distance with NIRSpec at R~300 at a single band. The figure shows how the SNR varies over the
wavelength region (1 – 5µm). Each simulation was done using 25 transit observations.
Simulations were done for super-Earths in 1.5-day orbits, which places little constraint on
the number of transit observations based on JWST’s five-year mission lifetime. According to
Kopparapu et al. 2013, super-Earths at the inner edge of the habitable zone of a 0.2 M! M-dwarf
will be in ~20-day orbits and those at the outer edge will be in ~45 day orbits. Therefore, twenty-
five transits will even be obtainable for super-Earths in the habitable zones of their M-dwarf
parent stars. Furthermore, based on JWST’s five-year mission, a super-Earth in the outer edge of
the habitable zone of a .2 M! M-dwarf could be observed with up to 40 transits.
Returning to Figure 1, the most striking difference lies between the SNRs of the H-poor and
H-rich simulations (right side vs. left side – note the difference in scale). The highest SNR of the
H-poor simulations (~14 for a 10 M⊕, 1000 K planet) is no greater than the lowest SNR of the H-
rich simulations (~14 for a 1 R⊕, 400 K planet). This is an effect of the atmosphere’s scale
height, which is inversely proportional to the mean molecular weight of the gas. Planets with
hydrogen-rich atmospheres have larger scale heights because of the small molecular weight.
Such atmospheres are normally associated with gas giant planets that have no visible solid
surface. Were we to assume an Earth with a hydrogen-rich atmosphere (that would, in fact, not
be stable for long), the resulting S/N would be lower; likewise for an H-rich atmosphere that is
cloudy.
Figures 2 and 3 illustrate simulated spectra for a range of temperatures and masses based on
the SNR calculations above. Figure 2 shows NIRSpec observations for H-rich atmosphere
models at 1 – 5µm assuming 25 transits as a function of temperature and mass. Figure 3 displays
the same information as Figure 2 but for H-poor atmosphere models. Each of these plots is
organized in a grid where each panel shows the output JWST simulated spectrum at a given
planetary temperature (x-axis) and mass (y-axis). The red curve is the Howe & Burrows model
as published; the black curve is the simulated spectrum at a distance of 4.5 pc. We also display a
spectrum in green, which represents the furthest distance NIRSpec could observe the object and
still maintain a SNR of at least ~15 (our threshold for detection; shown by the green arrow).
Looking at the case of the 10 M⊕, 1000 K, H-rich planet in the upper right hand panel of
Figure 2, there is clear detection of H2O and CH4 at all the wavelengths mentioned above at 4.5
pc (black curve). At 50 pc (green curve) the H2O/CH4 absorption feature at 1.4 µm and the H2O
absorption feature at 1.8 µm are much harder to resolve even though the red Howe & Burrows
model makes it seem otherwise. Extrapolating this case out to the other masses and temperatures,
it can be deduced that with 25 transits, an H-rich planet could be observed at M ≥ 1 M⊕ and T ≥
400 K out to 11 pc. For planets with larger masses and higher temperatures it would be possible
to observe a H-rich planet beyond 50 pc if it had a M ≥ 10 M⊕ and/or T ≥ 1000 K.
The H-poor case presents very different results. Looking at the case of the 10 M⊕, 1000 K,
H-poor planet in the upper right hand panel of Figure 3, there is still clear detection of H2O and
CH4 at all the wavelengths mentioned above at 4.5 pc (black curve). It is not possible to observe
this planet out to 50 pc, as with the H-rich case. It is only possible to obtain a SNR of 15 with 25
transits at 26 pc. The boxes shaded in red are cases where an SNR of 15 was not achievable.
Therefore, the spectrum of any H-poor planet with T ≤ 400 K would be dominated by noise.
Only if the H-poor planet has T > 700 K, will it be possible to resolve any of the water bands
especially at M ≤ 4 M⊕.
0.9989
0.9988
)
)
)
⊕
M
0.9987
0
1
)
0.9986
)
)
)
)
)
)
)
)
)
)
)
)
)
)
)
)
)
0.9985
9.9965•10-1
9.9960•10-1
)
4.5$pc,$V=9$
22$pc,$V=17$
2
3
))
9.9955•10-1
⊕
M
9.9950•10-1
4
9.9945•10-1
)
)
)
)
)
)
)
)
)
)
4.5$pc$
14$pc,$V=15$
3
2
9.9940•10-1
9.9935•10-1
1.0000•100
)
9.9995•10-1
)
)
9.9990•10-1
⊕
M
1
9.9985•10-1
)
)
)
)
)
)
)
)
9.9980•10-1
)
)
s
s
a
M
9.9975•10-1
Temperature):))))400)K
2
4.5$pc$
11$pc,$V=14$
3
Dec 2013
4.5$pc$
50$pc,$V=21$
2
3
4
4.5$pc$
43$pc,$V=20$
2
3
4
4.5$pc$
37$pc,$V=19$
2
3
4
4.5$pc$
2
29$pc,$V=18$
4
3
0.9990
0.9988
0.9986
0.9984
0.9982
0.9993
0.9992
0.9991
0.9990
0.9989
0.9988
0.9987
0.9998
0.9996
0.9994
0.9992
4.5$pc$
25$pc,$V=18$
)
2
)))))))))))700)K)
3
4
)
)
0.9990
)
4.5$pc$
40$pc,$V=20$
)
2
))))))))1000)K)
3
4
0.9989
0.9988
0.9987
0.9986
0.9985
0.9984
0.9983
0.9997
0.9996
0.9995
0.9994
0.9993
1.0001
1.0000
0.9999
0.9998
0.9997
0.9996
0.9995
)
4
4
4
)
Figure 2. Hydrogen Rich Atmosphere- Simulated JWST/NIRSpec spectra at three different
temperatures (400, 700, and 100 K) and three different masses (1, 4, and 10 M⊕). All simulations
assume 25 transits and each panel span 1 – 5 µm in wavelength. The simulations include effects
of pointing jitter, flat field errors and pixel response function. Model spectra were taken from
Howe & Burrows (2012) for 3x solar metallicity, assuming a parent star with similar properties
to GJ1214 (V=14.67) except with varying distances from 4.5 – 50 pc. The spectra in black
illustrate the most optimistic case at 4.5 pc, whereas the green curve shows the furthest distance
you can achieve SNR ~ 15. For this case of a H-rich atmosphere, JWST/NIRSpec can provide
excellent characterization of water and methane absorption features at 2.7 and 4.6 µm out to 11
pc at M ≥ 1 M⊕ and T ≥ 400 K.
9.9878•10-1
%
%
%
9.9876•10-1
%
⊕
9.9874•10-1
M
0
1
9.9872•10-1
%
%
%
%
%
%
%
%
%
%
%
%
9.9870•10-1
9.9868•10-1
9.9962•10-1
1
%
9.9960•10-1
%
%
9.9958•10-1
⊕
M
9.9956•10-1
4
%
9.9878•10-1
9.9876•10-1
9.9874•10-1
9.9872•10-1
9.9870•10-1
9.9868•10-1
3$pc,$V=7$
2
3
4
9.9866•10-1
9.9960•10-1
5
1
Dec 2013
9.9880•10-1
9.9875•10-1
9.9870•10-1
9.9865•10-1
4.5$pc$,V=9$
9.2$pc,$V=16$
2
3
4.5$pc$
9.9860•10-1
9.9965•10-11
5
4
26$pc,$V=18$
2
3
4
5
9.9958•10-1
9.9956•10-1
9.9954•10-1
9.9952•10-1
9.9960•10-1
9.9955•10-1
9.9950•10-1
3$pc$
2
3
9.9950•10-1
9.9994•10-1
1
5
4
3$pc$
2
3
9.9945•10-1
9.9994•10-11
5
4
4.5$pc$
7.7$pc,$V=12$
2
3
4
5
9.9954•10-1
%
%
%
%
%
%
%
%
%
%
%
%
%
%
9.9952•10-1
9.9950•10-1
9.9996•10-1
1
%
9.9994•10-1
%
⊕
M
9.9992•10-1
1
%
%
%
%
:
s
s
9.9990•10-1
a
M
9.9988•10-1
1
9.9992•10-1
9.9990•10-1
9.9988•10-1
9.9992•10-1
9.9990•10-1
9.9988•10-1
9.9986•10-1
1
5
%
4
%
3$pc$
%
2
%%%%%%%%%%700%K%
3
4
%
9.9986•10-1
5
1
%
%
3$pc$
%
2
3
%%%1000%K%
4
5
3$pc$
Temperature%:%%%%400%K
2
3
Figure 3. Hydrogen Poor Atmosphere- Simulated JWST/NIRSpec spectra at three different
temperatures (400, 700, and 100 K) and three different masses (1, 4, and 10 M⊕). All
simulations assume 25 transits and each panel span 1 – 5 µm in wavelength. The simulations
include effects of pointing jitter, flat field errors and pixel response function. Model spectra (red)
were taken from Howe & Burrows (2012) for pure water atmospheres, assuming a parent star
with similar properties to GJ1214 (V=14.67) except with varying distances from 3 – 26 pc. The
spectra in black illustrate the most optimistic case at 4.5 or 3 pc, whereas the green curve shows
the furthest distance you can achieve SNR ~ 15. For this case of a H-poor atmosphere,
JWST/NIRSpec will be limited to high-signal to noise characterization of more massive and/or
hotter super-Earths (yellow and white boxes). The red shaded boxes are cases where it is not
possible to achieve SNR of 15, even at 3 pc.
5. Summary and Future Work
JWST will be astronomy’s most space powerful telescope to date, and it is poised to
transform our understanding of the Universe. In this white paper, we provided realistic
simulations of JWST to characterize the increasing population of nearby super-Earth exoplanets
orbiting M dwarfs.
Our results are based on simulations derived from a JWST simulator combined with the latest
exoplanet transit models from Howe & Burrows (2012). We executed a small sample of
Dec 2013
simulations of primary transits, studied the changes in SNR, and interpolated these results over a
larger parameter space (mass, temperature, and distance). Based on a detectability criterion of an
SNR of ~15, we determine that 25 transits on JWST/NIRSpec is sufficient to detect water and
methane bands at 2.7 and 4.6 µm out to 11 pc at M ≥ 1 M⊕ and T ≥ 400 K. For the H-poor case
at 25 transit observations, a temperature above 700 K will be necessary to resolve any water
bands, especially for planets with M < 4 M⊕.
Following the promising results from this project, we will expand our study to repeat these
calculations with the NIRISS instrument on JWST, to expand the set of host stars to G and K
spectral types, and to include additional atmosphere models (e.g., intermediate between H-poor
and H-rich). Modeling of MIRI spectra from primary transits, and examination of secondary
transit spectra, constitute a separate future effort.
We thank Avi Mandell and Mark McCaughrean for helpful conversations, comments, and
suggestions.
References
Anglada-Escudé, G., Arriagada, P., Vogt, S. S., et al. 2012, ApJ, 751, L16
Batalha, Natalie M., Kepler Team, 2013, ApJ, 204, 24
Charbonneau, D., Berta, K., Irwin, J., et al. 2009, Nature, 462, 891
Clampin, M. Comparative Planetology, JWST Whitepaper 2010
Crouzet, N., McCullouhj, P. R., Burke, C., Long, D. 2012, ApJ, 761, 7
Deming, D., Harrington, J., Lauglin, G., 2007, ApJ, 667, L199
García-Muñoz, A., Zapatero Osorio, M.R., Barrena, R., Montañés-Rodríguez, P., Martín, E.L.
and Pallé, E.2012. ApJ. 755, 103.
Kaltenegger, L., Udry, S., Pepe, F., 2011, eprint arXiv: 1108.3561
Kopparapu, R., Ramirez, R., Kasting, J., et al. 2013, ApJ, 765, 131
Mayor, M., Udry, S., Lobis, C., et al. 2009, A&A, 493, 639
Miller-Ricci, E., Seager, S., Sasselov, D., 2009, ApJ, 690, 1056
Pepe, F., Lovis, C., Ségransan, D., et al. 2011, A&A, 534, A58
Ricker, G. R., Latham, D. W., Vanderspek, R. K., et al. 2010 AAS 21545006
Sing, D. K., Point, F., Aigrain, S., et al. 2011, MNRAS, 416, 1443
Tuomi, M., Jones, H. R. A., Jenkins, J. S., et al. 2013, A&A, 551, A79
Vogt, S. S., Butler, R. P., Rivera, E. J., et al. 2010, ApJ, 723, 954
|
1102.3209 | 1 | 1102 | 2011-02-16T00:12:21 | Clouds and the Faint Young Sun Paradox | [
"astro-ph.EP",
"physics.ao-ph"
] | We investigate the role which clouds could play in resolving the Faint Young Sun Paradox (FYSP). Lower solar luminosity in the past means that less energy was absorbed on Earth (a forcing of -50 Wm-2 during the late Archean), but geological evidence points to the Earth being at least as warm as it is today, with only very occasional glaciations. We perform radiative calculations on a single global mean atmospheric column. We select a nominal set of three layered, randomly overlapping clouds, which are both consistent with observed cloud climatologies and reproduce the observed global mean energy budget of Earth. By varying the fraction, thickness, height and particle size of these clouds we conduct a wide exploration of how changed clouds could affect climate, thus constraining how clouds could contribute to resolving the FYSP. Low clouds reflect sunlight but have little greenhouse effect. Removing them entirely gives a~forcing of +25 Wm-2 whilst more modest reduction in their efficacy gives a forcing of +10 to +15 Wm-2. For high clouds, the greenhouse effect dominates. It is possible to generate +50 Wm-2 forcing from enhancing these, but this requires making them 3.5 times thicker and 14 K colder than the standard high cloud in our nominal set and expanding their coverage to 100% of the sky. Such changes are not credible. More plausible changes would generate no more that +15 Wm-2 forcing. Thus neither fewer low clouds nor more high clouds can provide enough forcing to resolve the FYSP. Decreased surface albedo can contribute no more than +5 Wm-2 forcing. Some models which have been applied to the FYSP do not include clouds at all. These overestimate the forcing due to increased CO2 by 20 to 25% when CO2 is 0.01 to 0.1 bar. | astro-ph.EP | astro-ph | Manuscript prepared for Clim. Past
with version 3.2 of the LATEX class copernicus.cls.
Date: 5 July 2021
1
1
0
2
b
e
F
6
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
0
2
3
.
2
0
1
1
:
v
i
X
r
a
Clouds and the Faint Young Sun Paradox
C. Goldblatt1,* and K. J. Zahnle1
1Space Science and Astrobiology Division, NASA Ames Research Center, MS 245-3,
Moffett Field, CA 94035, USA
*now at: Astronomy Department, University of Washington, Box 351580, Seattle, WA 98195, USA
Abstract. We investigate the role which clouds could play in
resolving the Faint Young Sun Paradox (FYSP). Lower solar
luminosity in the past means that less energy was absorbed
on Earth (a forcing of −50 W m−2 during the late Archean),
but geological evidence points to the Earth being at least as
warm as it is today, with only very occasional glaciations.
We perform radiative calculations on a single global mean
atmospheric column. We select a nominal set of three lay-
ered, randomly overlapping clouds, which are both consis-
tent with observed cloud climatologies and reproduce the ob-
served global mean energy budget of Earth. By varying the
fraction, thickness, height and particle size of these clouds
we conduct a wide exploration of how changed clouds could
affect climate, thus constraining how clouds could contribute
to resolving the FYSP. Low clouds reflect sunlight but have
little greenhouse effect. Removing them entirely gives a forc-
ing of +25 W m−2 whilst more modest reduction in their effi-
cacy gives a forcing of +10 to +15 W m−2. For high clouds,
the greenhouse effect dominates. It is possible to generate
+50 W m−2 forcing from enhancing these, but this requires
making them 3.5 times thicker and 14 K colder than the stan-
dard high cloud in our nominal set and expanding their cover-
age to 100% of the sky. Such changes are not credible. More
plausible changes would generate no more that +15 W m−2
forcing. Thus neither fewer low clouds nor more high clouds
can provide enough forcing to resolve the FYSP. Decreased
surface albedo can contribute no more than +5 W m−2 forc-
ing. Some models which have been applied to the FYSP do
not include clouds at all. These overestimate the forcing due
to increased CO2 by 20 to 25% when pCO2 is 0.01 to 0.1 bar.
1
Introduction
Earth received considerably less energy from the Sun early
in Earth's history than today; ca. 2.5 Ga (billion years be-
Correspondence to: C. Goldblatt ([email protected])
fore present) the sun was only 80% as bright as today. Yet
the geological evidence suggests generally warm conditions
with only occasional glaciation. This apparent contradiction
is known as the Faint Young Sun Paradox (FYSP, Ringwood,
1961; Sagan and Mullen, 1972). A warm or temperate cli-
mate under a faint sun implies that Earth had either a stronger
greenhouse effect or a lower planetary albedo in the past, or
both.
In this study, we focus on the role of clouds in the
FYSP. We both examine how their representation in models
affects calculations of changes in the greenhouse effect, and
constrain the direct contribution that changing clouds could
make to resolving the FYSP.
Clouds have two contrasting radiative effects. In the spec-
tral region of solar radiation (shortwave hereafter) clouds
are highly reflective. Hence clouds contribute a large part
of Earth's planetary albedo (specifically the Bond albedo,
which refers to the fraction of incident sunlight of all wave-
lengths reflected by the planet).
In the spectral region of
terrestrial thermal radiation (longwave hereafter), clouds are
a strong radiative absorber, contributing significantly to the
greenhouse effect. Cloud absorption is largely independent
of wavelength (they approximate to "grey" absorbers), in
contrast to gaseous absorbers which absorb only in certain
spectral regions corresponding to the vibration -- rotation lines
of the molecules.
Despite the obvious, first-order, importance of clouds in
climate, it has become conventional to omit them in mod-
els of early Earth climate and use instead an artificially high
surface albedo. As described by Kasting et al. (1984):
Clouds are not included explicitly in the model;
however, their effect on the radiation balance is
accounted for by adjusting the effective albedo to
yield a mean surface temperature of 288 K for the
present Earth. The albedo is then held fixed for
all calculations at reduced solar fluxes. ...we feel
that the assumption of constant albedo is as good
2
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
as can be done, given the large uncertainties in the
effect of cloud and ice albedo feedbacks.
In effect, the surface is whitewashed in lieu of putting clouds
in the atmosphere. This assumption has been used exten-
sively in the models from Jim Kasting's group (Kasting et al.,
1984; Kasting and Ackerman, 1986; Kasting, 1987, 1988;
Kasting et al., 1993; Pavlov et al., 2000, 2003; Kasting and
Howard, 2006; Haqq-Misra et al., 2008), which, together
with parametrisations and results based on these models (for
example, Kasting et al., 1988; Caldeira and Kasting, 1992a,b;
Kasting et al., 2001; Kasting, 2005; Tajika, 2003; Bendt-
sen and Bjerrum, 2002; Lenton, 2000; Franck et al., 1998,
2000; von Bloh et al., 2003a,b; Lenton and von Bloh, 2001;
Bergman et al., 2004) have dominated early Earth palaeocli-
mate and other long term climate change research for the last
two and a half decades. The validity of this method has not
previously been tested.
Whilst Kasting's approach is that changes to clouds are so
difficult to constrain that one cannot justifiably invoke them
to resolve the FYSP, others are more bold. Some recent pa-
pers have proposed cloud-based resolutions to the FYSP.
Rondanelli and Lindzen (2010) focus on increasing the
warming effect of high clouds, finding that a total covering
of high clouds which have been optimised for their warm-
ing effect could give a late Archean global mean temper-
ature at freezing without increasing greenhouse gases. To
justify such extensive clouds, they invoke the "iris" hypoth-
esis (Lindzen et al., 2001) which postulates that cirrus cov-
erage should increase if surface temperatures decrease (this
hypothesis has received much criticism, e.g. Hartmann and
Michelsen, 2002; Chambers et al., 2002).
Rosing et al. (2010) focus on decreasing the reflectivity of
low level clouds so that the Earth absorbs more solar radia-
tion. To justify this, they suggest that there was no emission
of the important biogenic cloud condensation nuclei (CCN)
precursor dimethyl sulphide (DMS) during the Archean and,
consequently, clouds were thinner and had larger particle
sizes.
We note that both Rondanelli and Lindzen (2010) and Ros-
ing et al. (2010) predict early Earth temperatures substan-
tially below today's, which we do not consider a satisfactory
resolution of the FYSP.
Shaviv (2003) and Svensmark (2007) proposes less low-
level cloud on early Earth due to fewer galactic cosmic
rays being incident on the lower troposphere. The under-
lying hypothesis is of a correlation between galactic cos-
mic ray incidence and stratus amount, through CCN cre-
ation due to tropospheric ionization (Svensmark and Friis-
Christensen, 1997; Svensmark, 2007). This has received ex-
tensive study in relation to contemporary climate change and
has been refuted (e.g. Sun and Bradley, 2002; Lockwood and
Frohlich, 2007; Kristj´ansson et al., 2008; Bailer-Jones, 2009;
Calogovic et al., 2010; Kulmala et al., 2010, and references
therein).
In this study, we comprehensively asses how the radia-
tive properties of clouds, and changes to these, can affect the
FYSP. First, we explicitly evaluate how accurate cloud-free
calculations of changes in the greenhouse effect are with re-
spect to atmospheres with clouds included. We do this by
considering a very wide range of cloud properties within
a single global mean atmospheric column, finding a case
study which matches Earth's energy budget, then comparing
the effect of more greenhouse gas in this column to a cloud-
free calculation. Second, we conduct a very wide exploration
of how changing clouds could directly influence climate. We
vary fraction, thickness, height and particle size of the clouds
and vary surface albedo. We do not advocate any particular
set of changes to clouds. Rather, constraints on what direct
contribution clouds can make towards resolving the FYSP
emerges from our wide exploration of the phase space.
The heyday of clouds in 1-D models was the 1970s and
1980s. Improvements in radiative transfer codes and compu-
tational power over the last 30 years allow us to contribute
new insight to the problem, in particular by widely exploring
phase-space. Nonetheless, these classic papers retain their
relevence and are instructive as to how one might treat clouds
in such simple models (e.g. Schneider, 1972; Reck, 1979;
Wang and Stone, 1980; Stephens and Webster, 1981; Char-
lock, 1982). With specific relevance to the FYSP, Kiehl and
Dickinson (1987) included clouds in their model of methane
and carbon dioxide warming on early Earth, and calculated
radiative forcings from some changed cloud cases (our re-
sults here agree with this older work). Rossow et al. (1982)
considered cloud feedbacks for early Earth.
Regarding whether a cloud-free model will correctly cal-
culate the increased greenhouse effect with increases gaseous
absorbers, we hypothesise that it will lead to an overestima-
tion in the efficacy of enhanced greenhouse gases.
In the
absence of clouds, the broadest range of absorption is due to
water vapour. However, whilst water vapour absorbs strongly
at shorter and longer wavelengths, it absorbs weakly between
8 and 15 µm. This region of weak absorption is known as the
water vapour window. It is coincident with the Wein peak of
Earth's surface thermal emission at 10 µm. Thus the water
vapour window permits a great deal of surface radiation to
escape to space unhindered. Other greenhouse gases -- and
clouds -- do absorb here, so are especially important to the
greenhouse effect. With clouds absorbing some fraction of
the radiation at all wavelengths, the increase in absorption
with increased greenhouse gas concentration would be less
than if clouds were absent. Therefore, we think that a cloud-
free model would overestimate increased gaseous absorp-
tion with increased greenhouse gas abundance and underes-
timate the greenhouse gas concentrations required to keep
early Earth warm.
Comparison of cloudy and cloud-free radiative forcings in
the context of anthropogenic climate change (Pinnock et al.,
1995; Myhre and Stordal, 1997; Jain et al., 2000) supports
our hypothesis. For CO2, a clear-sky calculation overesti-
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
3
mates the radiative forcing by 14%. For more exotic green-
house gases, which are optically thin at standard conditions
(CFCs, CCs, HCFCs, HFCs, PFCs, bromocarbons, iodocar-
bons), clear-sky calculations overestimate radiative forcing
by 26 -- 35%. CH4 and N2O are intermediate; their clear-sky
radiative forcings are overestimated by 29% and 25%, re-
spectively (Jain et al., 2000).
A roadmap of our paper is as follows. In Sect. 2 we de-
scribe our general methods, verification of the radiative trans-
fer scheme and the atmospheric profile we use. In Sect. 3
we deal specifically with the development of a case study of
three cloud layers representing the present climate and the
model sensitivity to this. In Sect. 4 we compare cloudy and
cloud-free calculations of the forcing from increased green-
house gas concentration. In Sect. 5 we explore what direct
forcing clouds could impart, and in section 6 we evaluate
the aforementioned cloud-based hypotheses for resolving the
FYSP.
2 Methods
2.1 Overview
Using a freely available radiative transfer code, we develop
a set of cloud profiles for single-column models which is in
agreement both with cloud climatology and the global mean
energy budget. This serves as the basis for comparison of
radiative forcing with a clear sky model and with changed
cloud properties.
2.2 Radiative forcing
In work on contemporary climatic change, extensive use is
made of radiative forcing to compare the efficacy of green-
house gases (e.g. Forster et al., 2007). This is defined as
the change in the net flux at the tropopause with a change in
greenhouse gas concentration, calculated either on a single
fixed temperature -- pressure profile or a set of fixed profiles
and in the absence of climate feedbacks. Surface tempera-
ture change is directly proportional to radiative forcing, with
a radiative forcing of approximately 5 W m−2 being required
to cause a surface temperature change of 1 K (see Fig. 7 of
Goldblatt et al., 2009b). Note that the tropopause must be
defined as the level at which radiative heating becomes the
dominant diabatic heating term (Forster et al., 1997), i.e. the
lowest level at which the atmosphere is in radiative equilib-
rium.
We base all our analyses on radiative forcings here. As
we make millions of radiative transfer code evaluations, sav-
ing in computational cost from comparing radiative forcings
rather than running a radiative-convective climate model is
significant, and facilitates the wide range of comparisons pre-
sented.
Table 1. GAM profile at levels (layer boundaries). Note that the
tropopause is at 100 hPa.
Pressure
(Pa)
Altitude
(km)
10
20
30
50
100
200
300
500
1000
2000
3000
5000
10000
15000
20000
25000
30000
35000
40000
45000
50000
55000
60000
65000
70000
75000
80000
85000
90000
95000
100000
64.739
59.912
56.951
53.114
47.763
42.393
39.339
35.612
30.842
26.290
23.671
20.445
16.204
13.727
11.914
10.461
9.237
8.168
7.215
6.352
5.562
4.834
4.159
3.529
2.938
2.381
1.855
1.356
0.882
0.431
0.000
Temperature Water vapour
(K)
230.00
245.61
252.88
260.00
266.29
260.33
254.31
243.24
228.11
222.10
218.71
212.59
206.89
211.83
219.01
225.87
233.27
240.52
247.19
253.27
258.62
263.15
267.14
270.73
274.00
277.05
279.84
282.28
284.08
285.85
289.00
(g/kg)
0.0036
0.0036
0.0036
0.0036
0.0035
0.0033
0.0032
0.0032
0.0031
0.0030
0.0029
0.0026
0.0023
0.0048
0.0153
0.0456
0.1852
0.3751
0.6046
0.8866
1.2365
1.6525
2.1423
2.7049
3.3366
4.1602
5.2152
6.3997
7.8771
9.5702
11.1811
Ozone
(ppmv)
1.080
1.384
1.626
1.974
2.600
5.484
6.810
7.242
7.490
6.169
4.780
2.250
0.516
0.344
0.160
0.122
0.089
0.070
0.058
0.051
0.047
0.045
0.045
0.044
0.042
0.039
0.035
0.033
0.032
0.031
0.031
2.3 Global Annual Mean atmosphere
We perform all our radiative transfer calculations on a sin-
gle Global Annual Mean (GAM) atmospheric profile (Ta-
ble 1). This is based on the GAM profile of Christidis et al.
(1997) with some additional high altitude data from Jain
et al. (2000). Surface albedo is set as 0.125 (Trenberth et al.,
2009). For standard conditions we use year 2000 gas compo-
sitions: 369 ppmv CO2, 1760 ppbv CH4 and, 316 ppbv N2O.
We use present day oxygen and ozone compositions through-
out the work. Whilst comparisons without these might be
interesting, they would necessitate using a different temper-
ature profile in order to be self consistent. This would signif-
icantly complicate our methods, so no such calculations are
performed. For solar calculations, we use the present solar
flux and a zenith angle of 60◦.
Calculating radiative forcings on a single profile does in-
troduce some error relative to using a set of profiles for var-
ious latitudes (Myhre and Stordal, 1997; Freckleton et al.,
1998; Jain et al., 2000). However, as this is a methodological
4
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
paper concerning single column radiative-convective models,
it is the appropriate approach to take here.
2.4 Radiative transfer code and verification
We use the Atmosphere Environment Research (AER) Rapid
Radiative Transfer Model (RRTM, Mlawer et al., 1997;
Clough et al., 2005), longwave version 3.0 and shortwave
version 2.5, which are available from http://rtweb.aer.com
(despite different version numbers, these were both the most
recent versions at the time of the research). RRTM has been
parameterised for pressures between 0.01 and 1050 hPa and
for temperatures deviating no more than 30 K from the stan-
dard mid-latitude summer (MLS) profile. We have verified
that the GAM profile we use is within this region of pressure-
temperature space. The cloud parameterisations in RRTM
which we select follow Hu and Stamnes (1993) for water
clouds and Fu et al. (1998) for ice clouds.
RRTM has been designed primarily for contemporary at-
mospheric composition. Our intended use is for different
atmospheric composition (higher greenhouse gas concentra-
tions), so it is necessary for us to independently test the
performance of the model under these conditions (Collins
et al., 2006; Goldblatt et al., 2009b). Following the ap-
proach of Goldblatt et al. (2009b) we directly compare long-
wave clear sky radiative forcings from RRTM to the AER
Line-by-Line Radiative Transfer Model (LBLRTM, Clough
et al., 2005). These runs are done on a standard Mid-Latitude
Summer (MLS) profile (McClatchey et al., 1971; Anderson
et al., 1986) to take advantage of the large number of com-
putationally expensive LBLRTM runs performed by Gold-
blatt et al. (2009b). Performance of the codes is evaluated
at three levels: the top of the atmosphere (TOA), the MLS
tropopause at 200 hPa and the surface. Upward and down-
ward fluxes are considered separately. The surface is taken
to be a black body, so the upward flux depends only on tem-
↑
perature (F
lw,surf = σT 4∗ ). The downward longwave flux at
the TOA is zero. Neither vary with greenhouse gas concen-
trations, so changes in the net flux at these levels depends on
one radiation stream only. At the tropopause the net flux is
the sum of the two streams. It is defined positive downwards,
Flw = F
↓
lw− F
↑
lw .
(1)
In addition to the radiative flux, we show (Fig. 1) the forcing
Flw = Flw− Flw,std ,
where Flw,std is the flux at preindustrial conditions and the
flux gradient (change of flux with changing gas concentra-
tion)
(2)
∂F
∂X
≈ ∆F
∆X
=
Fi+1− Fi
Xi+1− Xi
,
(3)
where Fi is the flux at gas concentration Xi (Goldblatt et al.,
2009b).
Our focus is on comparison of cloudy to cloud-free pro-
files within RRTM, so we do not need high accuracy cal-
culations of early Earth radiative forcings. We can there-
fore use rather relaxed and qualitative thresholds for ac-
ceptable model performance relative to LBLRTM: we re-
quire continuous and monotonic response to changing green-
house gas concentration (no saturation), the forcing should
be smooth and monotonic and divergence from the LBLRTM
flux gradient should be limited. For CO2, RRTM forcing
is not smooth or monotonic below pCO2=10−4 bar so this
region is excluded (see Fig. 1). CO2 concentrations up to
pCO2=10−1 bar are used, though there is some underestima-
tion of radiative forcing by RRTM above pCO2=10−2 bar.
Also, collision-induced absorption (absorption due to for-
bidden transitions) becomes important at pCO2 ∼ 0.1 bar (J.
Kasting, private communication) but coefficients for these
are not included in the HITRAN database on which both
RRTM and LBLRTM absorption coefficients are based.
Therefore, it is emphasised that the radiative forcings pre-
sented here for high CO2 will be underestimates, but valid
for intra-comparison.
The comparison of RRTM to LBLRTM (Fig. 1) is only for
the purpose of validating clear sky radiative forcing in the
context of this methodological study. We have undoubtedly
used RRTM outside its design range. This is not intended as
an assessment of its use for the contemporary atmosphere or
for anthropogenic climate change.
3 Cloud representation and model tuning
3.1 Practical problems and observational guidance
Generation of an appropriate cloud climatology for this work
is not straightforward. Two fundamental problems are short-
comings in available cloud climatologies and averaging to
a single profile. Concerning climatologies, the problem is
one of observations: surface observers will see the lowest
level cloud only, satellites will see the highest level of cloud
only. Radiosondes are cloud penetrating and cloud proper-
ties may be inferred from measured relative humidity, but the
spatial and temporal coverage of radiosonde stations is lim-
ited. See Wang et al. (2000) and Rossow et al. (2005) for ex-
tensive discussion of what progress can be made. Similarly,
radar can profile clouds, but such observations are sparse.
Concerning averaging, the dependence of the global energy
budget on cloud properties is expected to be non-linear: one
should not expect that a linear average of global cloud phys-
ical properties would translate into a set of clouds whose ra-
diative properties would give energy balance in a single co-
lumn. Nonetheless, available temporally and spatially aver-
aged data for cloud properties can guide how clouds should
be represented in the model.
Rossow et al. (2005) deduce zonally-averaged cloud frac-
tion profiles using a combination of International Satellite
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
5
Fig. 1. RRTM performance for CO2 for each flux and each level. Colours (online only) and markers are: black + for LBLRTM, magenta
× for RRTM. Shaded areas are range from Quaternary minimum (180 ppmv) to SRES maximum (1248 ppmv) concentration. Grey lines in
these areas are solid for pre-industrial (287 ppmv) and dashed for year 2000 (369 ppmv) concentrations.
Cloud Climatology Program (ISCCP) and radiosonde data
(Fig. 2). The existence of three distinct cloud layers and the
pressure levels of these are immediately apparent when av-
eraging the data meridionally (Fig. 3). Following Rossow
and Schiffer (1999), we divide the clouds into three groups,
with divisions at 450 and 700 hPa. "Low" cloud coresponds
to cumulus, stratocumulus and stratus clouds. "Mid" level
clouds correspond to altocumulus, altostratus and nimbo-
stratus. "High" clouds correspond to cirrus and cirrostra-
tus. Absolute cloud fractions cannot be extracted directly
from these data as information on how the clouds overlap
is lost in temporal and spatial averaging. A simple ap-
proach to give indicative values is to assume either maxi-
mum or random overlap within each group (high, middle and
low), then to scale these cloud amounts by a constant such
that randomly overlapping the three groups gives the IPCC
mean global cloud fraction of 67.6% (Rossow and Schif-
fer, 1999). Maximum and random overlap within groups
give cloud fractions [fhigh,fmid,flow]=[0.24,0.25,0.43] and
[fhigh,fmid,flow]=[0.25,0.29,0.39], respectively.
Averaged cloud optical thickness or water paths are more
difficult
to constrain, as they are not directly available
from the Rossow et al. (2005) data set (W. Rossow, per-
sonal communication, 2009). We proceed with ISCCP data
only. Rossow and Schiffer (1999) report water paths of
[Whigh,Wmid,Wlow]=[23,60,51] g m−2.
ISCCP data are
Fig. 2. Average cloud fraction with altitude following Rossow et al.
(2005, and W. Rossow, personal communication, 2009), for January
and July, land and ocean. White areas are where there is either no
land (the Southern and Arctic Oceans) or no ocean (Antarctica).
Flux (Wm−2)F↓ at Surface340350360370380390F↓ at 200 hPa1020304050F at 200 hPa−290−280−270−260−250−240−230−220−210−200−190F↑ at 200 hPa230240250260270280290300310F↑ at TOA240250260270280290300Forcing (Wm−2)−1001020304050−1001020−30−20−10010203040506070−60−50−40−30−20−1001020−40−30−20−1001020pCO2(bar)Flux gradient (Wm−2bar−1)10−510−410−310−210−1102103104105106pCO2(bar)10−510−410−310−210−1101102103104105106pCO2(bar)10−510−410−310−210−1102103104105106pCO2(bar)10−510−410−310−210−1−106−105−104−103−102pCO2(bar)10−510−410−310−210−1−106−105−104−103−102Pressure (hPa)Land: Jan 020040060080010000102030405060708090100sin(latitude)Pressure (hPa)Land: Jul −1−0.500.51020040060080010000102030405060708090100Ocean: Jan 0102030405060708090100sin(latitude)Ocean: Jul −1−0.500.5101020304050607080901006
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
Table 2. Parameter values used in the large cloud tuning ensemble.
Optical depth depends logarithmically on water path. Water path
and cloud fraction are varied independently for each layer. Water
paths range from optically thin to optically thick clouds (Curry and
Webster, 1999) with 10 values. Cloud fractions range from 5% to
100% coverage with 20 cases. Effective radius is for water clouds
(low and mid level) and generalised effective size is for ice clouds
(high). There there are 103×203=8×106 cases in total.
Fixed properties
High
Mid
Cloud top (hPa)
Cloud base (hPa)
Liquid or ice
Effective radius (µm)
Generalised effective size (µm)
300
350
Ice
--
75
550
650
Liquid
11
--
Low
750
900
Liquid
11
--
Variable properties
Water path (g m−2)
Cloud fraction
All layers
[100.4,100.6,100.8,...,102.2]
[0.05, 0.10, 0.15, . . . , 1.00]
Even with the assumption that each cloud is homoge-
neous, each of our three cloud layers is represented by
a cloud base and top, water path,
liquid:ice ratio, and
effective particle sizes for liquid and ice particles, giv-
ing 6 degrees of freedom for each cloud. With three lay-
ers, there are eight permutations for overlap, contributing
another 7 degrees of freedom for the fractional coverage.
A total of 25 degrees of freedom is clearly impossible to
explore fully. As a necessary simplification, we fix the
cloud base and top, take clouds to be either liquid (low
and mid clouds) or ice (high clouds) and fix the particle
size (following Rossow and Schiffer, 1999). We assume that
cloud layers are randomly overlapped, so each cloud layer
can be represented by a single fraction from which the over-
lap is calculated (many GCMs use a "maximum-random"
overlap method where cloud fractions in adjacent layers are
correlated; this is not relevant here as our discrete cloud lay-
ers are separated by intervals of clear sky, e.g. see Hogan and
Illingworth, 2000).
Random overlap is easiest to explain for the case of two
cloud levels (A and B), with cloud fractions a and b. Fraction
ab of the sky would have both cloud layers, fraction a(1−b)
would only have level A clouds and fraction (1−a)b would
only have level B clouds, and fraction (1−a)(1−b) would be
cloud free. With three cloud layers, we have eight columns.
Each column is evaluated separately in both longwave and
shortwave spectral regions and the final single column is
found as a weighted sum of these 16 evaluations. Different
cloud fractions can be accounted for in this summation, re-
ducing the number of RRTM evaluations needed.
For each cloud layer, cloud fraction and water path are
varied widely whilst the other four parameters are fixed (Ta-
Fig. 3. Average cloud fraction with altitude following Rossow et al.
(2005, and W. Rossow, personal communication, 2009). (a) North-
ern Hemisphere. (b) Southern Hemisphere and (c) global for Jan-
uary (green), July (purple) and mean (black). Grey horizontal lines
separate high, mid- and low-level clouds.
from downward looking satellite data only and overlap is not
accounted for. Whilst the low cloud value will indicate low
clouds only, high and mid level cloud values may include
opacity contributions from the lower clouds which they ob-
scure (see Fig. 2). Hence these water paths are indicative
only.
Using fine resolution spatially and temporally resolved
data would help resolve these issues. However, to do so
would be beyond the scope of this work and, we feel, it is
beyond what is necessary to address the first-order questions
which are the subject of this paper.
3.2 Development of cloud profiles
We need to develop a set of cloud profiles which appropri-
ately represents Earth's cloud and energy budget climatolo-
gies. By necessity, we shall need to simplify cloud proper-
ties, tune our model clouds and consider sensitivity of the
model energy budget to these clouds.
Cloud fraction (%)Pressure (hPa)N. hem.0510152025303505001000Cloud fraction (%)Pressure (hPa)S. hem.0510152025303505001000Cloud fraction (%)Pressure (hPa)Global0510152025303505001000C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
7
ble 2). In all of the resulting cloud cases, we run the radia-
tive transfer code for both standard and elevated CO2 levels
(369 ppmv and 50 000 ppmv), giving 16 million runs in total.
We refer to model runs in which we include clouds in this
way as including "real clouds". This is meant in the sense
that clouds are included in the radiative transfer code in a
detailed and physically based manner. This is by contrast to
previous models (e.g. Kasting et al., 1984), where clouds are
represented non-physically by changing surface albedo.
3.3 Sensitivity experiment
For each cloud case that we have defined (the large ensemble,
Table 2), we calculate the radiative forcing at the tropopause
(Ftrop), to which change in surface temperature is propor-
tional. Radiative forcing is the change in net flux (here with
increasing CO2):
Ftrop = F[trop,highCO2]− F[trop,stdCO2] ,
where F in each case is the net flux as a sum of longwave
and shortwave spectral regions and upward and downward
streams of radiation:
(4)
F = (F
↓
lw− F
↑
lw) + (F ↓
sw− F ↑
sw).
(5)
We consider two subsets of the large ensemble:
1. Cloud sets which give energy balance at the TOA. This
is the most basic constraint on a possible climate. With
FTOA<5 W m−2, a subset of 1.0 million cases re-
mains. A relatively large FTOA,stdCO2 is allowed as
variations in the water path are coarse, but it is corrected
for by calculating radiative forcings so cannot bias the
outcome.
at
2. Cloud sets which give
energy balance
the
TOA and are close to observed longwave and
the TOA (Trenberth et al.,
shortwave fluxes at
Constraints are FTOA,stdCO2<5 W m−2,
2009).
↑TOA,SW,stdCO2 <115 W m−2
and
95<F
227<F ↑TOA,LW,stdCO2 <247 W m−2).
This gives
a subset of 36 985 cases.
Fig. 4. Histogram of radiative forcings from two subsets of cloud
profiles. Cloud sets which give energy balance at the TOA (sub-
set 1) in light grey, cloud sets which give energy balance at the
TOA and are close to observed longwave and shortwave fluxes at
the TOA superimposed darker (green online). Dashed vertical line
(red online) shows radiative forcing cloud-free case for comparison.
Table 3. Cloud properties used in case study. ftotal=0.66.
Property
High
Mid
Low
Cloud top (hPa)
Cloud base (hPa)
Cloud fraction
Water path (g m−2)
Liquid or ice
Generalised effective size ( µm)
Effective radius (µm)
250
300
0.25
20
Ice
75
--
500
600
0.25
25
Liquid
--
11
700
850
0.40
40
Liquid
--
11
The distribution of radiative forcings in these two sub-
sets is shown relative to the cloud-free radiative forcing of
41.3 W m−2 (Fig. 4). The maximum radiative forcing from
subset 1 is 40.2 W m−2; all physically plausible cloud sets
give a smaller radiative forcing than a cloud-free model. Sub-
set 2 -- of cloud sets which give Earth-like climate -- has
a mean radiative forcing of 34.6 with a standard deviation of
1.3 W m−2. The radiative forcing from the cloud-free case
is 4.9 standard deviations above the mean radiative forcing
from realistic, Earth-like, clouds.
Note that we perform these runs at the standard solar con-
stant, as for all model runs herein. Given that CO2 is not a
strong absorber in the shortwave spectral region, selecting a
lower solar constant for the sensitivity experiment would not
cause any noticeable change to Fig. 4. For example, using a
solar flux 80% of the present value yields forcings different
by 0.2 to 0.3 W m−2.
3.4 Case study selection
As discussed, there are many problems associated with se-
lecting a set of cloud profiles. However, the radiative forc-
ings from CO2 enhancement in all Earth-like cloud sets are
closely grouped (Fig. 4) and the mean of these is signifi-
cantly different from the cloud-free case. This justifies defi-
nition of a case study which can be used to represent Earth's
clouds. To do this from subset 2, we additionally constrain
Radiative forcing (W m−2)Frequency1520253035404500.511.522.533.544.55x 1048
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
Fig. 5. Comparison of global annual mean energy budgets: two estimates for contemporary climate, (a) Trenberth et al. (2009), based on
a composite of data and (b) Zhang et al. (2004), from ISCCP-FD data, compared to models used in this paper, (c) case study with real clouds
(d) cloud-free model.
cloud fractions (each layer and the resultant total) and water
paths of each layer to be close to climatological values, op-
timising for agreement with longwave and shortwave fluxes
at the TOA. We found that, whilst shortwave fluxes could
be found that were in good agreement with climatological
values, the outgoing longwave fluxes were slightly too high
in all cases from ensemble 2. Increasing the height of the
clouds by 50 hPa gives a better fit for longwave fluxes. Case
study cloud properties are given in Table 3 and the radiative
outcome in Fig. 5c.
3.5 Cloud-free case
In order to compare calculated cloudy and cloud-free radia-
tive forcings we need a cloud-free model as a comparison
case. To generate this, we follow Kasting et al. (1984) and
tune the surface albedo of the GAM profile to achieve energy
balance at the top of the atmosphere for a clear sky profile.
The required surface albedo is 0.264.
4 Real clouds and cloud-free model compared
First, consider the energy budget at standard conditions rel-
ative to observational climatology (Fig. 5). Our case study
with real clouds (Fig. 5c) is in very good agreement with ob-
servational climatology (Fig. 5a,b). By contrast, almost all
of the variable fluxes in the cloud-free model (Fig. 5d) are
markedly different; omitting clouds means that the global
energy budget is not properly represented. Overall in the
cloud-free model, more absorption of solar radiation (only
81 W m−2 of outgoing shortwave radiation is reflected rather
than 106 W m−2, a lower overall planetary albedo) is bal-
anced by a weaker greenhouse effect (with an elevated out-
going longwave flux of 261 W m−2 rather than 236 W m−2
and depressed downward longwave at the surface).
Whilst no 1-D model can perfectly represent global cli-
mate, our real cloud case study, which is constrained by ob-
servational cloud climatology, gives good agreement with the
observed energy budget. This justifies using it as an internal
standard, against which the cloud-free model can be com-
pared.
Again at standard conditions, compare the spectrally re-
solved fluxes (Fig. 6). In the shortwave, the difference in ad-
sorption between cloud-free and real cloud models (Fig. 6e)
has the same shape as the Planck function of solar radiation
(Fig. 6c). This is because the surface albedo is constant with
wavelength by definition and the wavelength dependence of
cloud scattering is weak. Rayleigh scattering is spectrally de-
pendent (short wavelengths are preferentially scattered), but
this is a small term (14.5 W m−2 in the cloud-free case). By
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
9
SW for both cases in black, F ↑TOA
Fig. 6. Comparison of spectrally resolved energy budgets in
Real Cloud (RC, blue) and Cloud Free (CF, red) models. Black
lines are for both cases. Green is for differences (CF−RC).
SW in colours. (b) F ↑surf
(a) F ↓TOA
LW
for both cases in black, F ↑TOA
(c) Absorp-
tion of solar radiation: A=F ↓TOA
SW (d) Greenhouse
effect: G=F ↑surf
LW (e) Difference in solar absorp-
tion: DA=A(CF)−A(RC) (f) Difference in greenhouse effect:
DG=G(CF)−G(RC).
LW in colours.
SW −F ↑TOA
LW −F ↑TOA
contrast, in the longwave, there is strong spectral dependence
in the differences between the real cloud and cloud-free mod-
els. The cloud-free model has a weaker greenhouse effect
than real clouds in the water vapour window region. Whilst
other spectral regions are optically thick (with gaseous ab-
sorption by water vapour and carbon dioxide dominating) the
water vapour window is optically thin and the cloud green-
house is important.
Now consider the effect of changing CO2 concentration
(Fig. 7). Radiative forcing is strongly overestimated by the
cloud-free model relative to our real cloud case study; to pro-
duce a given radiative forcing, twice as much CO2 is needed
with the real cloud case study than is indicated by the cloud-
free model.
The radiative forcing in the longwave is an order of magni-
tude larger than the radiative forcing in the shortwave region,
so we focus on the longwave region when comparing spec-
trally resolved forcing (Fig. 8). The greenhouse effect with
real clouds is stronger at standard conditions than the cloud-
free model (inclusion of cloud greenhouse). However, the
greenhouse forcing (increase in strength of the greenhouse
effect), is larger in the cloud-free model. This is true across
all bands where CO2 imparts a greenhouse effect and is most
important in the water vapour window. Here, the atmosphere
Fig. 7. Radiative forcing with increasing pCO2. Real clouds in blue
and cloud-free in red. Present pCO2 marked (.).
LW in black, F ↑TOA
Fig. 8. Comparison of spectrally resolved longwave forcings for in-
crease from standard to 50 000 ppmv CO2 in real cloud (RC, blue)
and cloud-free (CF, red) models. Green is for differences in cases
(CF−RC) and black is for fluxes common between cases. (a) RC:
F ↑surf
LW dashed blue for standard CO2 and solid
blue for elevated CO2; (b) CF: F ↑surf
LW dashed red
for standard CO2 and solid red for elevated CO2; (c) greenhouse
forcing from increased CO2: G=G(HighCO2)−G(StdCO2)
where G=F ↑surf
LW ; (d) difference in greenhouse forcing:
DG=G(CF)−G(RC).
LW in black, F ↑TOA
LW −F ↑TOA
Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)01230200400600Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)01230100200300400Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)0123−2002040Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)01020300102030Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)0102030051015Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)0102030−6−4−20acebdf10−410−310−210−1−100102030405060pCO2 (bar)Radiative forcing (Wm−2)Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)01020300102030Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)01020300102030Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)0102030−50510Wavelength, λ (µm)Flux, Fλ (W m−2 µm−1)010203000.511.52acbd10
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
Fig. 9. Absorption cross sections for CO2 (green) and water vapour (purple) from HITRAN (shown at 900 hPa and 285 K). Horizontal lines
indicate the cross section for which the gas has an optical depth of unity, solid purple for the GAM water vapour column, dashed green for
various CO2 concentrations. The column depth of the atmosphere is 2.1×1025 molecules cm−2.
is optically thin in the absence of clouds, so the effect of in-
creasing CO2 is large even though its absorption lines are
weak (Fig. 9). With clouds, these regions will be optically
thicker initially so increasing CO2 has less of an effect.
At 15 µm,
increasing CO2 causes increased longwave
emission. This is due to increased emission in the strato-
sphere and is therefore unaffected by tropospheric clouds.
Our GAM profile includes O3 which absorbs at 9.5 µm and
9.7 µm. This would be absent in the anoxic Archean atmo-
sphere, making the water vapour window optically thinner.
The overestimation of forcing by the cloud-free model is,
therefore, likely larger than suggested here and even more
CO2 would actually be needed to cause equivalent warming.
The other perturbation to consider is the change in in-
coming solar flux. The cloud-free model absorbs a higher
proportion of the incoming solar flux (has a lower plane-
tary albedo), so will have a proportionately larger response
to changing solar flux. For an 20% decrease in solar flux,
representative of the late Archean, the decrease in absorbed
solar flux is 52.3 W m−2 for the cloud-free model compared
to 47.2 W m−2 for the real-cloud case study. Taking the
Archean to be lower solar flux but higher CO2, this error
is of opposite sign to the error in radiative forcing from in-
creased CO2 and around half the magnitude. Whilst these
errors could be said to partially offset in these conditions,
reliance on errors of opposing sign is not strong.
5 Variation of cloud and surface properties
The problem of cloud feedback on climate change is noto-
riously difficult. We do not attempt to address this in full;
rather, we explore how variations in cloud amounts and prop-
erties could affect climate. In all cases here, our baseline case
is the real cloud case study and we consider the radiative ef-
fect of changes in cloud or surface properties. As comparison
values, if we increase or decrease the humidity in the model
profile by 10% (50%), the radiative forcings are 1.7 W m−2
(7.8 W m−2) and −1.9 W m−2 (−11.2 W m−2), respectively.
A fainter sun in the late Archean is equivalent to a forcing of
around −50 W m−2 (assuming a planetary albedo of 0.3).
5.1 Surface albedo
In the cloud-free model, the use of a non-physical surface
albedo means that real changes in surface albedo cannot be
considered. This limitation is removed with explicit clouds.
We limit discussion here to changes in surface albedo not
from ice, though the use of a physically realistic surface with
CO2 = 500 ppmvCO2 = 50000 ppmvCO2 = 5000 ppmvWater vapour columnC. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
11
flection) but a positive forcing in the longwave (greenhouse
effect). The greenhouse effect operates by absorption of ther-
mal radiation emitted by a warm surface followed by emis-
sion at a lower temperature. Therefore the magnitude of
changes in the greenhouse effect varies with cloud height,
as higher clouds are colder. For low and mid level clouds,
shortwave effects dominate and increasing cloud fraction
or thickness will cause a net negative forcing (cooling the
planet). For high clouds, shortwave and longwave effects are
of similar magnitudes so the character of the net response is
more complicated. For water paths less than 350 g m−2, high
clouds cause a net positive forcing (greenhouse warming the
planet). The converse is true above 350 g m−2, but such high
water paths would typically correspond to deep convective
clouds, not high clouds (cirrus or cirrostratus) (Rossow and
Schiffer, 1999). Positive forcing is maximum for ∼70 g m−2
high clouds.
5.3 Cloud particle size
Cloud particle size depends very strongly on the availabil-
ity of cloud condensation nuclei (CCN). Whilst the global
mean droplet size is 11 µm, this is biased by smaller droplets
over land (average 8.5 µm), where there are more CCN than
over the ocean (average 12.5 µm). Over the ocean, around
half of CCNs are presently derived from oxidation products
of biogenic dimethyl sulphide (DMS), especially sulphuric
acid (there are various oxidation pathways of DMS (e.g.
von Glasow and Crutzen, 2004), but only sulphuric acid can
cause nucleation of new droplets (Kreidenweis and Seinfeld,
1988)). The climatic feedbacks involving DMS (Charlson
et al., 1987) have been subject of long debate. Whilst DMS is
prevalent today due to production by eukaryotes, other bio-
genic sulphur gases are produced by bacteria, in particular
hydrogen sulphide (H2S) and methyl mercaptan (CH3SH)
(Kettle et al., 2001). These will react chemically to form sul-
phates, which will provide CCN.
We do not delve deeply into CCN feedbacks here, but
accept that various changes in the Earth system (e.g. atmo-
spheric oxidation state, sulphur cycle, volcanic fluxes, bi-
ological fluxes) may well have changed CCN availability.
Fewer CCN give larger cloud drops, which should both rain
out quicker (so less cloud) and be less reflective. Conversely,
more CCN give more extensive and more reflective clouds.
We consider the effect of changing liquid droplet size
by factors of 0.5, 1.5 and 2 relative to the case study
(reff =11 µm) and ice particle size by factors of 0.5, 1.5 and
1.87 relative to the case study (DGE=75 µm; the maximum
of the parameterisation used is 140 µm).
In Fig. 12, we
show the net (shortwave plus longwave) radiative forcing
from changing particle size for all water paths and fractions.
The effect is strongest for low clouds. With no change to
cloud fraction or water path, increasing reff by 50% gives
a forcing of 7.5 W m−2 and doubling reff gives a forcing
Fig. 10. Radiative forcing with changed surface albedo. (.) is the
case study and ( ◦ ) is the end-member case of an ocean covered
planet.
RC means that a parameterised ice-albedo feedback could be
included in 1-D climate models, a significant improvement
on the status quo.
The surface albedo we use of 0.125 represents a weighted
average of land (0.214) and ocean (0.090) albedos (Trenberth
et al., 2009). Continental volume is generally thought to have
increased over time, with perhaps up to 5% of the present
amount at the beginning of the Archean and 20% -- 60% of the
present amount by the end of the Archean (e.g. Hawkesworth
and Kemp, 2006). In Fig. 10 we consider a range of varia-
tion of surface albedo appropriate for a changed land frac-
tion. For the end-member case relevant to the Archean of
a water-world, the radiative forcing is 4.8 W m−2. Without
land, relative humidity would likely be higher, contributing
extra forcing.
5.2 Cloud fraction and water path
There are more clouds over ocean than land (Fig. 2). The
zonally uninterrupted Southern Ocean is especially cloudy.
One might therefore expect that when there was less land
there would have been more cloud, and more still if there was
a greater extent of zonally uninterrupted ocean. Comparison
of the Northern and Southern Hemispheres of Earth (Fig. 3),
the former having a higher land fraction, may be indicative
of the minimum expected degree of variation. The Southern
Hemisphere has 20 -- 50% greater cloud fraction in each layer
than the Northern Hemisphere.
We consider a wide range of water paths, from optically
thin to thick clouds, and fractional cloud cover from zero to
1 for each cloud layer. In Fig. 11 we show the radiative forc-
ing from these clouds relative to no cloud in the given layer.
The competing shortwave and longwave effects of changing
clouds can readily be seen. Increasing fraction or water path
causes a negative forcing in the shortwave region (more re-
0.080.10.120.14−4−202468albedoRadiative forcing (Wm−2)12
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
Fig. 11. Cloud radiative forcing with cloud fraction and water path relative to no cloud in that layer. For each cloud layer, these properties are
varied whilst the clouds in other layers remain fixed at the case study values, marked (.). Particle sizes, ice/water ratio and height are as case
study. Colour/contour scale is in W m−2. For comparison, resolution of the late Archean FYSP would require a forcing of approximately
50 W m−2.
of 10.4 W m−2. Decreasing reff by 50% gives a forcing of
−13.6 W m−2.
Satellite observations of the modern ocean (Br´eon et al.,
2002) suggests a limit on how large droplet size actually be-
comes in nature. Particle size is rarely larger than 15 µm,
even in the remotest and least productive regions of the
ocean. Here, the DMS flux is low and remaining CCN derive
from abiological sources (e.g. sea spray). reff =15 µm can
then be seen as the baseline case for lower CCN availability,
corresponding to a 36% size increase relative to present day
mean (20% relative to present day ocean).
If there was a larger CCN flux, the droplet size for clouds
over land (reff =8.5 µm, 23% less than mean) is an indicator
of likely droplet size.
Larger droplets will rain out more effectively, but model
representations of this feedback vary dramatically (Pen-
ner et al., 2006; Kump and Pollard, 2008). For the case
of reff =17 µm droplets over the ocean, Kump and Pollard
(2008) choose a mid-strength assumption of this feedback,
implying a decrease of water path by a factor of 2.2. This
is marked (×) in the low cloud, 16.5 µm panel of Fig. 12;
the radiative forcing is then 15.4 W m−2, twice that of solely
increasing droplet size. Clearly, an increased precipitation
feedback is of first order importance and must be treated
carefully in any model addressing the climatic effect of
changed particle size.
5.4 Cloud height
To test the sensitivity to cloud height, each cloud layer is
raised or lowered 100 hPa and the forcing is calculated rel-
ative to the standard heights for all water paths and frac-
tions. As temperature decreases with height, higher clouds
emit at a lower temperature. It is this longwave effect which
is dominant. There are only small changes in the shortwave
effect, due to changed path length above the cloud (a greater
path length means decreased insolation due to more Rayleigh
scattering in the overlying atmosphere). In Fig. 13 we show
only the net forcing. For the high clouds in the case study,
which cover one-quarter of the sky and have a water path
−120−100−80−60−40−20High cloudWater path (g m−2)Shortwave10010110210320406080100Longwave−20020Net−160−140−120−100−80−60−40−20Mid cloudWater path (g m−2)1001011021032040−120−100−80−60−40−20−160−140−120−100−80−60−40−20Cloud fractionLow cloudWater path (g m−2)00.51100101102103Cloud fraction00.51−140−120−100−80−60−40−20Cloud fraction 00.51−150−100−50050100150C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
13
Fig. 12. Change in net cloud radiative forcing with cloud particle size, across a range of cloud fractions and water paths. Particle size is varied
for each layer independently (values in subplot titles), whilst all properties for other cloud layers remain as case study. The change is shown
relative to the case study (so panels for reff =11 µm and DGE=75 µm contain the same information as Fig. 11 net fluxes). Cloud fraction and
water path for case study are marked (.); values at the point of these markers are for changing particle size only, values elsewhere in each
panel are for changing water path or fraction too. Markers (×) and (+) refer to reduction in water path by factors of 2.2 and 3.7, respectively,
for comparison to Rosing et al. (2010), as discussed in the text. Marker (.) corresponds to the relatively thick and maximum extent clouds
invoked by Rondanelli and Lindzen (2010). Colour/contour scale is in W m−2.
of 20 g m−2, the effect of raising them is relatively small
(2.8 W m−2). There is a larger forcing from raising clouds
which are thicker or cover more of the sky initially; the
greater the radiative longwave effect of the cloud at its stan-
dard height, the greater the effect of changing it height would
be.
Changes in the temperature -- pressure structure of the at-
mosphere might have induced changes in clouds. The
Archean atmosphere was anoxic and did not have an ozone
layer (Kasting and Donahue, 1980; Goldblatt et al., 2006).
Consequently, there would likely not have been a strong
stratospheric temperature inversion, and deep atmospheric
convection may have reached higher altitudes, where the at-
mosphere is colder. A major source of high clouds is detrain-
ment of cirrus from deep convective clouds. Where detrain-
ment is due to wind sheer, this could then result in higher
clouds. Conversely, without an inversion a the tropopause,
cumulonimbus incus (anvil shaped clouds) will not form. As
the forcing from raising the high clouds in the case study is
small, other climatic effects might be larger (loss of ozone as
a greenhouse effect and lower stratospheric emission temper-
ature). Also, the pressure of Archean atmosphere was likely
not 1 bar. Not only was there no oxygen (0.21 bar today), but
the nitrogen inventory was likely different (Goldblatt et al.,
2009a). Varying pressure would have changed both the lapse
rate and tropopause pressure (Goldblatt et al., 2009a).
6 Evaluating cloud-based proposals to resolve the Faint
Young Sun Paradox
6.1
Increased cirrus
Rondanelli and Lindzen (2010) proposed that near total cov-
−40−200High cloudWater path (g m−2)DGE = 37.5 µm100101102103−120−100−80−60−40−200Mid cloudWater path (g m−2)reff = 5.5 µm100101102103−120−100−80−60−40−20020Cloud fractionLow cloudWater path (g m−2)reff = 5.5 µm 00.51100101102103−150−100−50050100150−20020DGE = 75 µm−100−80−60−40−200reff = 11 µm−120−100−80−60−40−20020Cloud fractionreff = 11 µm 00.51−150−100−50050100150020DGE = 112.5 µm−80−60−40−200reff = 16.5 µm−100−80−60−40−20020Cloud fractionreff = 16.5 µm 00.51−150−100−50050100150020DGE = 140 µm−80−60−40−200reff = 22 µm−80−60−40−20020Cloud fractionreff = 22 µm 00.51−150−100−5005010015014
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
cover the whole sky, not just one-quarter (similar to the sug-
gestion of Rondanelli and Lindzen, 2010). This gives a forc-
ing of 29.0 W m−2, insufficient to counter the ∼50 W m−2
deficit from the FYSP. If, in addition, we raise the cloud by
100 hPa (base at 200 hPa, making the cloud 14 K colder) the
total radiative forcing becomes 50.7 W m−2.
In principle, high clouds can resolve the FYSP. In prac-
tice, the requirement for total high level cloud cover seems
implausible and the requirement that the clouds are higher
(colder) is difficult to justify. That it takes an extreme end-
member case to provide only just enough forcing to resolve
the FYSP suggests that resolution with enhanced cirrus only
is not a strong hypothesis.
6.2 Decreased stratus
Rosing et al. (2010) propose that there were less CCN avail-
able in the Archean, due to lower DMS emissions prior to the
oxygenation of the atmosphere and widespread occurrence of
eukarya. They suggest an increase in droplet size from 12 µm
to 20 or 30 µm. Even over unproductive regions of today's
oceans, the effective radius of cloud particles rarely exceeds
15 µm (Br´eon et al., 2002), so it is difficult to see how such
large effective radii could be justified. Larger droplets lead
to more rain, so should make clouds thinner. To account for
this, Rosing et al. (2010) arbitrarily decrease the liquid wa-
ter path of their stratus clouds by a factor of 3.7, which is at
the high end of likely decreases (Penner et al., 2006). Even
with these very strong assumptions, their model temperature
is continually below the present temperature before 2 Ga.
In our framework of radiative forcings,
the effects of
changing effective radius and cloud water path are shown in
Fig. 12. For the strong but arguably plausible case (discussed
in Sect. 5.3) of doubling the effective radius and decreas-
ing water path by a factor of 2.2 gives a radiative forcing of
15.4 W m−2. For the yet stronger case of doubling the effec-
tive radius from 11 µm to 22 µm and decreasing cloud water
path by a factor of 3.7, the radiative forcing is 20.5 W m−2.
Removing low cloud entirely gives a forcing of 25.3 W m−2.
We therefore conclude that reducing stratus cannot by itself
resolve the FYSP.
A separate hypothesis (Shaviv, 2003; Svensmark, 2007)
proposes less stratus on early Earth due to fewer galactic
cosmic rays being incident on the lower troposphere. The
underlying hypothesis is of a correlation between galactic
cosmic ray incidence and stratus amount, through CCN cre-
ation due to tropospheric ionization (Svensmark and Friis-
Christensen, 1997; Svensmark, 2007). This hypothesis has
been refuted (e.g. Sun and Bradley, 2002; Lockwood and
Frohlich, 2007; Kristj´ansson et al., 2008; Bailer-Jones, 2009;
Calogovic et al., 2010; Kulmala et al., 2010, and references
therein): galactic cosmic rays cause the formation of at most
10% of CCNs and there is no correlation between galactic
cosmic ray incidence and cloudiness. Also of note is that
Shaviv (2003) requires a highly non-standard climate sen-
Fig. 13. Cloud radiative forcing with changed cloud height relative
to standard height clouds (see Fig. 11). Colour/contour scale is in
W m−2.
erage of cirrus clouds could resolve the FYSP. Their pro-
posed mechanism is that the planet would be colder and have
lower sea surface temperatures, which would give more cir-
rus coverage (the controversial "iris" hypothesis of Lindzen
et al., 2001), acting as a strong negative feedback on tem-
perature. The first premise here, of colder temperatures,
is contrary to the geological record; this suggests less fre-
quent glaciation through the Archean and Proterozoic than
in the Phanerozoic, not more. The second premise, of
strong cloud feedback, is based on a statistical relationship
for Earth's tropics (Lindzen et al., 2001) the authenticity of
which has been questioned (e.g. Hartmann and Michelsen,
2002; Chambers et al., 2002). Application to very cold tem-
peratures requires an extreme and unverifiable extrapolation.
Rondanelli and Lindzen (2010) describe the high level clouds
they use as "thin cirrus"; we note that the clouds they use ac-
tually have twice the water path of our standard high clouds.
In their sensitivity tests, using a thinner high level clouds
gives a weaker effect.
Here, we consider what would be required of cirrus or
other high level clouds for them to resolve the FYSP. In-
formed by the experiments above, we construct an optimum
cirrus cloud for warming: relative to our case study we make
it 3.5 times thicker (a water path of 70 g m−2) and make it
20100hPa higherCloud fraction00.51−200High cloudWater path (g m−2)100hPa lower1001011021030Mid cloudWater path (g m−2)1001011021030Cloud fractionLow cloudWater path (g m−2) 00.51100101102103−30−20−100102030C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
15
sitivity to force his model. In the sensitivity test where he
uses a more standard climate sensitivity, it results in a mean
temperature ∼ 0◦C during the Archean. Even if the underly-
ing hypothesis had not been refuted, the same arguments as
above would apply: plausible decreases in stratus are insuf-
ficient to resolve the FYSP.
7 Conclusions
When calculating radiative forcing from increased green-
house gas concentrations, we find that omitting clouds leads
to a systematic overestimate relative to models in which
clouds are included in a physically based manner. With
0.1 bar CO2 (the relevant quantity for a CO2 based resolu-
tion to the Faint Young Sun Paradox in the late Archean)
the overestimate of radiative forcing from modelling with-
out clouds approaches 10 W m−2 in our model, equivalent
to the clear-sky forcing from 100 ppm CH4. As the radia-
tive transfer code we use underestimates forcing from CO2 at
this level, and we include O3 in our profile, the difference be-
tween the real cloud case study and the cloud-free model here
must be seen as a lower bound on the error from omitting
clouds. For other greenhouse gases, especially those which
absorb strongly in the water vapour window, the overestima-
tion by a cloud-free model would likely be larger. This would
affect calculations of the warming by methane and ammonia
and of recently proposed Archean greenhouse gases, ethane
(Haqq-Misra et al., 2008) and OCS (Ueno et al., 2009).
The question of what direct effect clouds might have is
a more interesting and difficult one. We can address this best
by considering what radiative forcing can be generated in
both the shortwave and longwave spectral regions by chang-
ing cloud physical properties, and whether such changes in
cloud physical properties can be justified.
For solar radiation (shortwave), low level stratus clouds
have the greatest effect. Removing them from the model en-
tirely gives a forcing of 25 W m−2. Even this end-member
falls short of the 50 W m−2 that is needed to resolve the
FYSP. A more plausible combination of reduced fraction and
water path and increased droplet size would give a maximum
forcing of 10 -- 15 W m−2. However, suitable justification for
these changes does not come easily. Rosing et al. (2010)
asserted that DMS fluxes would be low in the Archean,
but there may well have been other biological and chemi-
cal sources for the sulphuric acid on which water condenses
(DMS is a precursor to this). For example, methyl mercap-
tan is produced abundantly by bacteria (Kettle et al., 2001).
Observations of clouds show that the effective radius rarely
becomes larger that 15 µm (Br´eon et al., 2002), which im-
plies that regionally low CCN flux does not lead to very large
droplets. If there were less land early in Earth's history and
more zonally uninterrupted ocean, one might expect there to
be more cloud rather than less (similar to how there is greater
cloud fraction in the Southern Hemisphere than the North-
ern Hemisphere). Also relating to low cloud, Shaviv (2003)
and Svensmark (2007) contend that less galactic cosmic rays
were incident on the troposphere during the Archean and this
would have led to less stratus. However, the underlying hy-
pothesis for this has been refuted.
For terrestrial radiation (longwave), high level clouds are
most important as they are coldest (the greenhouse effect de-
pends on the temperature difference between the surface and
the cloud). The end member case is 100% coverage of high
clouds which are optimised for their greenhouse effect, be-
ing both thicker and higher than our case study. Such an end
member case gives a forcing of 50 W m−2, which would just
be sufficient to resolve the FYSP. However, physical justifi-
cation for any of the required changes is lacking. Rondanelli
and Lindzen (2010) invoke a controversial negative feedback
of increased cirrus fraction with decreased temperature (the
"iris" hypothesis of Lindzen et al., 2001), but a true resolu-
tion to the FYSP should give temperatures equal or higher
than present. Thus, even if the "iris" hypothesis was cor-
rect, it would act to oppose warming. It is difficult to think
of other mechanisms to make high clouds wider and thicker.
Whether clouds should have been higher in the Archean may
warrant more study. The absence of the strong stratospheric
temperature inversion presently caused by ozone might con-
tribute. However, without increase in fraction or cloud water
path, the forcing will likely be less than 5 W m−2.
The question then naturally arises: How should one model
early Earth climate? Some would look first towards a gen-
eral circulation model (GCM), in order to better represent the
dynamics on which clouds depend. We disagree. Whilst dy-
namics are certainly important, it is unrealistic to think that in
the near future clouds could be resolved in a global scale cli-
mate model applicable to palaeoclimate. Even in "high reso-
lution" models used for anthropogenic global change, cloud
processes are parameterised sub-grid scale. As one moves to-
wards deep palaeoclimate research, one moves further from
the present atmospheric state for which the model may have
been designed and can be validated. A larger model therefore
introduces greater, and harder to track, uncertainty. Consid-
ering what radiative forcing or warming a given mixture of
greenhouse gases will impart is a first-order question, and
one which should be answerable with a first-order model.
A 1-D model is sufficient for this, but clouds must be in-
cluded. The appropriate starting point would likely be a
model with fixed cloud optical depth and fixed cloud top
temperature (see, for example Reck, 1979; Wang and Stone,
1980). For any proposed change to clouds, very great atten-
tion is needed to the feasibility of the mechanism involved.
To model these, one should probably look towards a cloud
microphysics resolving model, coupled to appropriate mod-
els of CCN supply and chemistry.
A stronger greenhouse effect likely contributes the largest
part of the forcing required to keep early Earth warm.
It
is important to remember, however, that the forcing from a
greenhouse gas depends on the logarithm of its abundance.
16
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
Thus, a modest forcing from clouds could have a large effect
on how much of a greenhouse gas in needed; indicatively,
a 10 W m−2 from CO2 requires a increase in concentration
of a factor of 2 to 3. A large atmospheric CO2 reservoir
(∼ 0.1 bar) may be slow to accumulate, as geochemical pro-
cesses (principally volcanic outgassing) contributing linearly
to concentration, so any other forcings may be rather useful
in resolving the FYSP.
In summary, it is necessary to include clouds in climate
models if these are to be accurate. Resolution of the faint
young sun paradox likely requires a combination of a few
different warming mechanisms, including strong contribu-
tions from one or more greenhouse gases. Changed clouds
could contribute warming, but this has yet to be justified --
and cooling caused by cloud changes is equally possible. Fu-
ture work will no doubt propose novel mechanisms to change
clouds. We hope that the results presented here will facil-
itate quick and accurate look-up the climatic effect of such
changes. Proposed cloud-based resolutions with only limited
greenhouse enhancement are not plausible.
Acknowledgements. Thanks to Richard Freedman for plotting the
HITRAN data for Fig. 9 and William Rossow for providing the data
for Fig. 2. Thanks to J. Kasting and I. Halevy for reviews and R.
Pierrehumbert for comments. C. G. was funded by a NASA Post-
doctoral Program fellowship. K. J. Z was funded by the NASA
Astrobiology Institute Ames Team and the NASA Exobiology pro-
gram.
References
Anderson, G. P., Clough, S. A., Kneizys, F. X., Chetwynd, J. H.,
and Shettle, E. P.: AFGL atmospheric constituent profiles (0 --
120km), Environmental Research Papers 954, Air Force Geo-
physical Laboratory, Hanscom AFB, MA, 1986.
Bailer-Jones, C. A. L.: The evidence for and against astronomical
impacts on climate change and mass extinctions: a review, Int.
J. Astrobiol., 8, 213 -- 239, doi:10.1017/S147355040999005X,
2009.
Bendtsen, J. and Bjerrum, J.: Vulnerability of climate on Earth to
sudden changes in insolation, Geophys. Res. Lett., 29, 1706, doi:
10.1029/2002GL014829, 2002.
Bergman, N. M., Lenton, T. M., and Watson, A. J.: COPSE: a new
model of biogeochemical cycling over the Phanerozoic, Am. J.
Sci., 304, 397 -- 437, 2004.
Br´eon, F.-M., Tanr´e, D., and Generoso, S.: Aerosol Effect on Cloud
Droplet Size Monitored from Satellite, Science, 295, 834 -- 838,
2002.
Caldeira, K. and Kasting, J. F.: The life span of the biosphere revis-
ited, Nature, 360, 721 -- 723, 1992a.
Caldeira, K. and Kasting, J. F.: Susceptibility of the Early earth to
irreversible glaciation caused by carbon-dioxide clouds, Nature,
359, 226 -- 228, 1992b.
Calogovic, J., Albert, C., Arnold, F., Beer, J., Desorgher, L.,
and Flueckiger, E. O.: Sudden cosmic ray decreases: No
change of global cloud cover, Geophys. Res. Lett., 37, L03 802,
doi:10.1029/2009GL041327, 2010.
Chambers, L., Lin, B., and Young, D.: Examination of new CERES
data for evidence of tropical Iris feedback, Journal of Climate,
15, 3719 -- 3726, 2002.
Charlock, T. P.: Cloud optical feedback and climate stability in a
radiative-convec tive model, Tellus, 34, 245 -- 254, 1982.
Charlson, R., Lovelock, J. E., Andreae, M., and Warren, S.: Oceanic
phytoplankton, atmospheric sulphur, cloud albedo and climate,
Nature, 326, 655 -- 661, doi:10.1038/326655A0, 1987.
Christidis, N., Hurley, M. D., Pinnock, S., Shine, K. P., and Walling-
ton, T. J.: Radiative forcing of climate change by CFC-11 and
possible CFC replacements, J. Geophys. Res., 102, 19 597 --
19 609, 1997.
Clough, S. A., Shephard, M. W., Mlawer, E. J., Delamere, J. S.,
Iacono, M. J., Cady-Pereira, K., Boukabara, S., and Brown,
P. D.: Atmospheric radiative transfer modeling: A summary of
the AER codes, J. Quant. Spectrosc. Ra., 91, 233 -- 244, 2005.
Collins, W. D., Ramaswamy, V., Schwarzkopf, M. D., Sun, Y., Port-
mann, R. W., Fu, Q., Casanova, S. E. B., Dufresne, J.-L., Fill-
more, D. W., Forster, P. M. D., Galin, V. Y., Gohar, L. K., In-
gram, W. J., Kratz, D. P., Lefebvre, M. ., Li, J., Marquet, P.,
Oinas, V., Tsushima, Y., Uchiyama, T., and Zhong, W. Y.: Radia-
tive forcing by well-mixed greenhouse gases: Estimates from cli-
mate models in the Intergovernmental Panel on Climate Change
(IPCC) Fourth Assessment Report (AR4), J. Geophys. Res., 111,
D14 317, doi:10.1029/2005JD006713, 2006.
Curry, J. A. and Webster, P. J.: Thermodynamics of the Atmo-
spheres and Ocean, Academic Press, London, 471 pp, 1999.
Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fa-
hey, D., Haywood, J., Lean, J., Lowe, D., Myhre, G., Nganga, J.,
Prinn, R., Raga, G., Schulz, M., and Van Dorland, R.: Changes
in Atmospheric Constituents and in Radiative Forcing, in: Cli-
mate Change 2007: The Physical Science Basis. Contribution of
Working Group I to the Fourth Assessment Report of the Inter-
governmental Panel on Climate Change, edited by Solomon, S.,
Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K., Tig-
nor, M., and Miller, H. L., Cambridge Univ. Press, Cantab. U.K.
and New York, NY, USA, 2007.
Forster, P. M. d. F., Freckleton, R. S., and Shine, K. P.: On aspects
of the concept of radiative forcing, Clim. Dynam., 13, 547 -- 560,
1997.
Franck, S., Kossacki, K. J., and Bounama, C.: Modelling the global
carbon cycle for the past and future evolution of the earth system,
Chem. Geol., 159, 305 -- 317, 1998.
Franck, S., Block, A., von Bloh, W., Bounama, C., Schellnhuber,
H. J., and Svirezhev, Y.: Reduction of biosphere life span as a
consequence of geodynamics, Tellus, 52B, 94 -- 107, 2000.
Freckleton, R. S., J., H. E., Shine, K. P., Wild, O., Law, K. S., and
Sanderson, M. G.: Greenhouse gas radiative forcing: effects of
averaging and inhomogeneities in trace gas distribution, Q. J. R.
Meteorol. Soc., 124, 2099 -- 2127, 1998.
Fu, Q., Yang, P., and Sun, W. B.: An accurate parameterization
of the intrared radiative properties of cirrus coluds for climate
models, J. Clim, 11, 2223 -- 2237, 1998.
Goldblatt, C., Lenton, T. M., and Watson, A. J.: Bistability of atmo-
spheric oxygen and the Great Oxidation, Nature, 443, 683 -- 686,
doi:10.1038/nature05169, 2006.
Goldblatt, C., Claire, M. W., Lenton, T. M., Matthews, A. J.,
Watson, A. J., and Zahnle, K. J.: Nitrogen-enhanced green-
house warming on early Earth, Nature Geosci., 2, 891 -- 896,
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
17
doi:10.1038/ngeo692, 2009a.
Goldblatt, C., Lenton, T. M., and Watson, A. J.: An evaluation
of the longwave radiative transfer code used in the Met Of-
fice Unified Model, Quart. J. Roy. Met. Soc., 135, 619 -- 633,
doi:10.1002/qj.403, 2009b.
Haqq-Misra, J. D., Domagal-Goldman, S. D., Kasting, P., and Kast-
ing, J. F.: A revised, hazy methane greenhouse for the Archean
Earth, Astrobiology, 8, 1127 -- 1137, doi:10.1089/ast.2007.0197,
2008.
Hartmann, D. L. and Michelsen, M.: No evidence for Iris, Bull.
Am. Met. Soc., 83, 249 -- 254, 2002.
Hawkesworth, C. J. and Kemp, A. I. S.: Evolution of the continental
crust, Nature, 443, 811 -- 817, 2006.
Hogan, R. J. and Illingworth, A. J.: Deriving cloud overlap statistics
from radar, Q. J. R. Meteorol. Soc., 126, 2903 -- 2909, 2000.
Hu, Y.-X. and Stamnes, K.: An Accurate Parameterization of the
Radiative Properties of Water Clouds Suitable for Use in Climate
Models, J. Clim., 6, 728 -- 742, 1993.
Jain, A. K., Briegleb, B. P., Minschwaner, K., and Wuebbles, D. J.:
Radiative forcings and global warming potentials of 39 green-
house gases, J. Geophys. Res., 105, 20 773 -- 20 790, 2000.
Kasting, J. F.: Theoretical constraints on oxygen and carbon-
dioxide concentrations in the Precambrian atmosphere, Precam-
brian Research, 34, 205 -- 229, 1987.
Kasting, J. F.: Runaway and moist greenhouse atmospheres and the
evolution of Earth and Venus, Icarus, 74, 472 -- 494, 1988.
Kasting, J. F.: Methane and climate during the Precambrian era,
Precambrian Res., 137, 119 -- 129, 2005.
Kasting, J. F. and Ackerman, T. P.: Climatic consequences of very
high-carbon dioxide levels in the Earth's early atmosphere, Sci-
ence, 234, 1383 -- 1385, 1986.
Kasting, J. F. and Donahue, T. M.: The Evolution of the Atmo-
spheric Ozone, J. Geophys. Res., 85, 3255 -- 3263, 1980.
Kasting, J. F. and Howard, M. T.: Atmospheric composition and
climate on the early Earth, Philos. T. Roy. Soc. B, 361, 1733 --
1741, 2006.
Kasting, J. F., Pollack, J. B., and Crisp, D.: Effects of high CO2
levels on surface temperature and atmospheric oxidation-state of
the early Earth, J. Atmos. Chem., 1, 403 -- 428, 1984.
Kasting, J. F., Toon, O. B., and Pollack, J. B.: How climate evolved
on the terrestrial planets, Scientific American, 258, 90 -- 97, 1988.
Kasting, J. F., Whitmere, D. P., and Reynolds, R. T.: Habitable
zones around main sequence stars, Icarus, 101, 108 -- 128, 1993.
Kasting, J. F., Pavlov, A. A., and Siefert, J. L.: A coupled
ecosystem -- climate model for prediction the methane concentra-
tion of in the Archean Atmosphere, Origins Life Evol. Biosphere,
31, 271 -- 285, 2001.
Kettle, A. J., Rhee, T. S., von Hobe, M., Poulton, A., Aiken, J., and
Andreae, M. O.: Assessing the flux of different volatile sulfur
gases from the ocean to the atmosphere, J. Geophys. Res., 106,
12 193 -- 12 210, doi:10.1029/2000JD900630, 2001.
Kiehl, J. T. and Dickinson, R. E.: A study of the radiative effects
fo enhanced atmospheric CO2 and CH4 on early Earth surface
temperatures, J. Geophys. Res., 92, 2991 -- 2998, 1987.
Kreidenweis, S. M. and Seinfeld, J. H.: Nucleation of sulfuric acid-
water and methanesulfonic acid-water solution particles: Impli-
cations for the atmospheric chemistry of organosulfur species,
Atmos. Env., 22, 283 -- 296, doi:10.1016/0004-6981(88)90034-
0, 1988.
Kristj´ansson, J. E., Stjern, C. W., Stordal, F., Fjaeraa, A. M., Myhre,
G., and J´onasson, K.: Cosmic rays, cloud condensation nuclei
and clouds -- a reassessment using MODIS data, Atmos. Chem.
Phys., 8, 7373 -- 7387, doi:10.5194/acp-8-7373-2008, 2008.
Kulmala, M., Riipinen, I., Nieminen, T., Hulkkonen, M., So-
gacheva, L., Manninen, H. E., Paasonen, P., Petj, T., Dal Maso,
M., Aalto, P. P., Viljanen, A., Usoskin, I., Vainio, R., Mirme,
S., Mirme, A., Minikin, A., Petzold, A., Hrrak, U., Pla-Dlmer,
C., Birmili, W., and Kerminen, V.-M.: Atmospheric data over
a solar cycle: no connection between galactic cosmic rays and
new particle formation, Atmos. Chem. Phys., 10, 1885 -- 1898,
doi:10.5194/acp-10-1885-2010, 2010.
Kump, L. R. and Pollard, D.: Amplification of Cretaceous
Warmth by Biological Cloud Feedbacks, Scinece, 320, 195,
doi:10.1126/science.1153883, 2008.
Lenton, T. M.: Land and ocean carbon cycle feedback effects on
global warming in a simple Earth system model, Tellus, 52B,
1159 -- 1188, 2000.
Lenton, T. M. and von Bloh, W.: Biotic feedback extends the life
span of the biosphere, Geophys. Res. Lett., 28, 1715 -- 1718, 2001.
Lindzen, R. S., Chou, M.-D., and Hou, A. Y.: Does the Earth have
an adaptive infrared iris?, Bull. Am. Met. Soc., 82, 417 -- 432,
2001.
Lockwood, M. and Frohlich, C.: Recent oppositely directed
trends in solar climate forcings and the global mean surface
air temperature, Proc. R. Soc. A, 463, 2447 -- 2460, doi:doi:
10.1098/rspa.2007.1880, 2007.
McClatchey, R. A., Fenn, R. W., Selby, J. E. A., Volz, F. E., and
Garling, J. S.: Optical properties of the atmosphere (revised),
Environmental Research Papers 354, Air Force Cambridge Re-
search Laboratories, 1971.
Mlawer, E. J., Taubman, S. J., Brown, P. D., Iacono, M. J., and
Clough, S. A.: Radiative transfer for inhomogeneous atmo-
spheres: RRTM, a validated correlated-k method for the long-
wave, J. Geophys. Res., 102, 16 663 -- 16 682, 1997.
Myhre, G. and Stordal, F.: Role of spatial and temporal variations in
the computaion of radiative forcing and GWP, J. Geophys. Res.,
102, 11 181 -- 11 200, 1997.
Pavlov, A. A., Kasting, J. F., Brown, L. L., Rages, K. A., and Freed-
man, R.: Greenhouse warming by CH4 in the atmosphere of
early Earth, J. Geophys. Res., 105, 11,981 -- 11,990, 2000.
Pavlov, A. A., Hurtgen, M. T., Kasting, J. F., and Arthur, M. A.:
Methane-rich Proterozoic atmosphere?, Geology, 31, 87 -- 90,
2003.
Penner, J. E., Quaas, J., Storelvmo, T., Takemura, T., Boucher, O.,
Guo, H., Kirkevag, A., Kristj´ansson, J. E., and Seland, O. .:
Model intercomparison of indirect aerosol effects, Atmos. Chem.
Phys., 6, 3391 -- 3405, http://www.atmos-chem-phys.net/6/3391/
2006/, 2006.
Pinnock, S., Hurley, M. D., Shine, K. P., Wallington, T. J., and
Smyth, T. J.: Radiative forcing of climate by hydrochlorofloro-
carbons and hydroflorocarbons, J. Geophys. Res., 100, 23 277 --
23 238, 1995.
Reck, R.: Comparison of fixed cloud-top temperature and fixed
cloud-top altitude approximations in the Manabe -- Wetherald
radiative-convective atmospheric model, Tellus, 31, 400 -- 405,
1979.
Ringwood, A.: Changes in solar luminosity and some possible ter-
restrial consequences, Geochimica et Cosmochimica Acta, 21,
18
C. Goldblatt and K. J. Zahnle: Clouds and the Faint Young Sun Paradox
0442(2000)013¡3041:CVSAIV¿2.0.CO;2, 2000.
Wang, W.-C. and Stone, P. H.: Effect of ice-albedo feedback on
global sensitivity in a one-dimensional radiative-convective cli-
mate model, J. Atmos. Sci., 37, 545 -- 552, 1980.
I.: Calculation of
Zhang, Y. C., Rossow, W. B., Lacis, A. A., Oinas, V., and
radiative fluxes from
Mishchenko, M.
the surface to top of atmosphere based on ISCCP and
other global data sets: Refinements of the radiative transfer
model and the input data, J. Geophys. Res., 109, D19 105,
doi:10.1029/2003JD004457, 2004.
295 -- 296, doi:10.1016/S0016-7037(61)80064-1, 1961.
Rondanelli, R. and Lindzen, R. S.: Can thin cirrus clouds in the
tropics provide a solution to the faint young Sun paradox?,
J. Geophys. Res., 115, D02 108, doi:10.1029/2009JD012050,
2010.
Rosing, M. T., Bird, D. K., Sleep, N. H., and Bjerrum, C. J.: No
climate paradox under the faint early Sun, Nature, 464, 744 -- 747,
doi:10.1038/nature08955, 2010.
Rossow, W. B. and Schiffer, R. A.: Advances in Understanding
Clouds from ISCCP, Bull. Am. Met. Soc., 80, 2261 -- 2287, 1999.
Rossow, W. B., Henderson-Sellers, A., and Weinreich, S. K.: Cloud
feedback: a stabilising effect for the early earth?, Science, 217,
1245 -- 1247, 1982.
Rossow, W. B., Zhang, Y., and Wang, J.: A Statistical Model
of Cloud Vertical Structure Based on Reconciling Cloud Layer
Amounts Inferred from Satellites and Radiosonde Humidity Pro-
files., J. Climate, 18, 3587 -- 3605, doi:10.1175/JCLI3479.1, 2005.
Sagan, C. and Mullen, G.: Earth and Mars: evolution of atmo-
spheres and surface temperatures, Science, 177, 52 -- 56, 1972.
Schneider, S. H.: Cloudiness as a global climatic feedback mech-
anism: the effects on the radiation balance and surface temper-
ature variations in cloudiness, J. Atmos. Sci., 29, 1413 -- 1422,
1972.
Shaviv, N. J.: Toward a solution to the early faint Sun paradox: A
lower cosmic ray flux from a stronger solar wind, J. Geophys.
Lett., 108, 1437, doi:10.1029/2003JA009997, 2003.
Stephens, G. L. and Webster, P. J.: Clouds and climate: sensitivity
of simple systems, J. Atmos. Sci., 38, 235 -- 247, 1981.
Sun, B. and Bradley, R. S.: Solar influences on cosmic rays and
cloud formation: A reassessment, J. Geophys. Res., 107, 4211,
doi:10.1029/2001JD000560, 2002.
Svensmark, H.: Cosmoclimatology: a new theory emerges, Astron.
Geophys., 48, 1.18 -- 1.24, 2007.
Svensmark, H. and Friis-Christensen, E.: Variations of cosmic ray
flux and global cloud coverage -- A missing link in solar-climate
relationships, J. Atmos. Sol. Terr. Phys., 59, 1225 -- 1232, 1997.
Tajika, E.: Faint young Sun and the carbon cycle: implications for
the Proterozoic global glaciations, Earth Planet. Sci. Lett., 214,
443 -- 453, 2003.
Trenberth, K. E., Fasullo, J. T., and Kiehl, J. T.: Earth's
global energy budget, Bull. Amer. Meteor. Soc., 90, 311 -- 324,
doi:10.1175/2008BAMS2634.1, 2009.
Ueno, Y., Johnson, M. S., Danielache, S. O., Eskebjerg, C. Pandey,
A., and Yoshida, N.: Geological sulfur isotopes indicate elevated
OCS in the Archean atmosphere, solving faint young sun para-
dox, Proc. Natl. Acad. Sci. U. S. A., 106, 14 784 -- 14 789, 2009.
von Bloh, W., Bounama, C., and Franck, S.: Cambrian Explosion
triggered by geosphere biosphere feedback, Geophys. Res. Lett.,
30, CLM 6 -- 1 -- CLM 6 -- 5, doi:10.1029/2003GL017928, 2003a.
von Bloh, W., Franck, S., Bounama, C., and Schellnhuber, H. J.:
Biogenic enhancement of weathering and the stability of the eco-
sphere, Geomicrobiol. J., 20, 501 -- 511, 2003b.
von Glasow, R. and Crutzen, P. J.: Model study of multiphase DMS
oxidation with a focus on halogens, Atmospheric Chemistry and
Physics, 4, 589 -- 608, doi:10.5194/acp-4-589-2004, http://www.
atmos-chem-phys.net/4/589/2004/, 2004.
Wang, J., Rossow, W. B., and Zhang, Y.: Cloud Vertical
Structure and Its Variations from a 20-Yr Global Rawin-
sonde Dataset, J.Climate, 13, 3041 -- 3056, doi:10.1175/1520-
|
1807.08209 | 2 | 1807 | 2018-08-19T19:48:50 | The HOSTS Survey for Exozodiacal Dust: Preliminary results and future prospects | [
"astro-ph.EP"
] | [abridged] The presence of large amounts of dust in the habitable zones of nearby stars is a significant obstacle for future exo-Earth imaging missions. We executed an N band nulling interferometric survey to determine the typical amount of such exozodiacal dust around a sample of nearby main sequence stars. The majority of our data have been analyzed and we present here an update of our ongoing work. We find seven new N band excesses in addition to the high confidence confirmation of three that were previously known. We find the first detections around Sun-like stars and around stars without previously known circumstellar dust. Our overall detection rate is 23%. The inferred occurrence rate is comparable for early type and Sun-like stars, but decreases from 71% [+11%/-20%] for stars with previously detected mid- to far-infrared excess to 11% [+9%/-4%] for stars without such excess, confirming earlier results at high confidence. For completed observations on individual stars, our sensitivity is five to ten times better than previous results. Assuming a lognormal luminosity function of the dust, we find upper limits on the median dust level around all stars without previously known mid to far infrared excess of 11.5 zodis at 95% confidence level. The corresponding upper limit for Sun-like stars is 16 zodis. An LBTI vetted target list of Sun-like stars for exo-Earth imaging would have a corresponding limit of 7.5 zodis. We provide important new insights into the occurrence rate and typical levels of habitable zone dust around main sequence stars. Exploiting the full range of capabilities of the LBTI provides a critical opportunity for the detailed characterization of a sample of exozodiacal dust disks to understand the origin, distribution, and properties of the dust. | astro-ph.EP | astro-ph | The HOSTS Survey for Exozodiacal Dust: Preliminary
results and future prospects
Ertel, S.a, Kennedy, G. M. b,c, Defr`ere, D.d, Hinz, P.a, Shannon, A. B.e,f, Mennesson, B.g,
Danchi, W. C.h, Gelino, C.g, Hill, J. M.i, Hoffmann, W. F.a, Rieke, G.a, Spalding, E.a,
Stone, J. M.∗a, Vaz, A.a, Weinberger, A. J.j, Willems, P.g, Absil, O.d, Arbo, P.a, Bailey, V. P.g,
Beichman, C.k, Bryden, G.g, Downey, E. C.a, Durney, O.a, Esposito, S.l, Gaspar, A.a,
Grenz, P.a, Haniff, C. A.m, Leisenring, J. M.a, Marion, L.d, McMahon, T. J.a,
Millan-Gabet, R.k, Montoya, M.a, Morzinski, K. M.a, Pinna, E.l, Power, J.i, Puglisi, A.l,
Roberge, A.h, Serabyn, E.g, Skemer, A. J.n, Stapelfeldt, K.g, Su, K. Y. L.a,
Vaitheeswaran, V.a, and Wyatt, M. C.o
aSteward Observatory, Department of Astronomy, University of Arizona, 993 N. Cherry Ave,
bDepartment of Physics, University of Warwick, Gibbet Hill Road, Coventry CV4 7AL, UK
cCentre for Exoplanets and Habitability, University of Warwick, Gibbet Hill Road, Coventry,
Tucson, AZ, 85721, USA
dSpace sciences, Technologies & Astrophysics Research (STAR) Institute, University of Li`ege,
CV4 7AL, UK
Li`ege, Belgium
eDepartment of Astronomy and Astrophysics, The Pennsylvania State University, State
fCenter for Exoplanets and Habitable Worlds, The Pennsylvania State University, State
College, PA 16801, USA
gJet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr., Pasadena,
College, PA 16802, USA
CA 91109, USA
hNASA Goddard Space Flight Center, Exoplanets & Stellar Astrophysics Laboratory, Code
iLarge Binocular Telescope Observatory, University of Arizona, 933 N. Cherry Avenue, Tucson,
667, Greenbelt, MD 20771, USA
jDepartment of Terrestrial Magnetism, Carnegie Institution of Washington, 5241 Broad
kNASA Exoplanet Science Institute, MS 100-22, California Institute of Technology, Pasadena,
Branch Road NW, Washington, DC, 20015, USA
AZ 85721, USA
lINAF-Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, I-50125 Firenze, Italy
mCavendish Laboratory, University of Cambridge, JJ Thomson Avenue, Cambridge CB3 0HE,
CA 91125, USA
nAstronomy Department, University of California Santa Cruz, 1156 High Street, Santa Cruz,
oInstitute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
CA 95064, USA
UK
∗ Hubble fellow.
Send correspondence to Steve Ertel, E-mail: [email protected]
8
1
0
2
g
u
A
9
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
9
0
2
8
0
.
7
0
8
1
:
v
i
X
r
a
ABSTRACT
The presence of large amounts of dust in the habitable zones of nearby stars is a significant obstacle for future
exo-Earth imaging missions. We executed the HOSTS (Hunt for Observable Signatures of Terrestrial Systems)
survey to determine the typical amount of such exozodiacal dust around a sample of nearby main sequence stars.
The majority of the data have been analyzed and we present here an update of our ongoing work. Nulling
interferometry in N band was used to suppress the bright stellar light and to detect faint, extended circumstellar
dust emission. We find seven new N band excesses in addition to the high confidence confirmation of three that
were previously known. We find the first detections around Sun-like stars and around stars without previously
known circumstellar dust. Our overall detection rate is 23%. The inferred occurrence rate is comparable for
early type and Sun-like stars, but decreases from 71+11−20% for stars with previously detected mid- to far-infrared
excess to 11+9−4% for stars without such excess, confirming earlier results at high confidence. For completed
observations on individual stars, our sensitivity is five to ten times better than previous results. Assuming a
lognormal luminosity function of the dust, we find upper limits on the median dust level around all stars without
previously known mid to far infrared excess of 11.5 zodis at 95% confidence level. The corresponding upper limit
for Sun-like stars is 16 zodis. An LBTI vetted target list of Sun-like stars for exo-Earth imaging would have a
corresponding limit of 7.5 zodis. We provide important new insights into the occurrence rate and typical levels
of habitable zone dust around main sequence stars. Exploiting the full range of capabilities of the LBTI provides
a critical opportunity for the detailed characterization of a sample of exozodiacal dust disks to understand the
origin, distribution, and properties of the dust.
Keywords: Exo-zodiacal dust, Interferometry, High contrast imaging, Habitable zone, Exo-Earth imaging,
Mid-infrared
1. INTRODUCTION
The presence of warm and hot dust in nearby planetary systems both provides important opportunities to study
their architecture and dynamics and impairs our ability to directly image and characterize Earth-like planets.
At temperatures between a few 100 K to ∼2000 K, such exozodiacal dust clouds (exozodis for short) are located
near the habitable zones (HZs, [1, 2, 3]) of their host stars and closer in [4, 5, 6, 7]. For a recent review see [8].
The dust may form from comet evaporation [9], local asteroid collisions (for young systems and at relatively
large separations from the star, [10]), and collisions of large bodies further out in the system from where it
migrates closer to the star due to Poynting-Robertson and stellar wind drag [11, 12]. It gravitationally interacts
with any planets in the system in addition to the star and is ground to (sub-)micron sizes through mutual
collisions of dust grains, after which the stellar radiation and wind pressure cause the grains to be expelled
from the system. The dust distribution may trace the location of the dust producing planetesimals or in case
of cometary origin the magnitude of the cometary activity. Structures in the dust distribution such as gaps and
resonant clumps can reveal the presence and parameters of planets [13, 14].
At the same time, bright dust clouds constitute the main source of confusion and noise for future exo-
Earth imaging missions. The mere presence of dust emission (scattered light or thermal emission may dominate
depending on the wavelength of observation) would add photon noise to the observations, even if the emission itself
could be perfectly removed. Moreover, dust clumps due to resonant trapping by a planet can be misinterpreted
as actual planets due to the limited angular resolution and signal-to-noise ratio of the observations. Thus,
the expected amount of dust present around the target stars determines the design of any mission to detect
exo-Earths, such as the required angular resolution and size of the primary mirror (e.g., [15, 16, 17]).
We have conducted the NASA funded HOSTS survey (Hunt for Observable Signatures of Terrestrial Systems)
to search for habitable zone dust around a sample of nearby stars and to measure their dust levels. We observed
in N band, where the flux ratio between warm (∼300 K) dust and the star is most favorable while the atmosphere
is still sufficiently transparent. The goal was to detect HZ dust levels only a few times more massive than the
Solar system level, corresponding to an N band excess emission of the order of 10−5. We thus used nulling
interferometry at the Large Binocular Telescope Interferometer (LBTI, [18]) in order to spatially disentangle the
dust emission from that of the star and to suppress the bright star light.
Figure 1. Illustration of the data obtained by nulling interferometry at the LBTI. Model observations are shown for two
example stars (top row : 1 L(cid:12), bottom row : 10 L(cid:12)) at a distance of 10 pc. The panels show from left to right: A model
image of a face-on disk, the LBTI transmission pattern, the transmission pattern applied to the model image, and the final
simulated observation after convolving with the single aperture PSF. The disk models with and without the transmission
pattern applied are shown in logarithmic scale, the other images are shown in linear scale. The dashed circle marks the
size of the HZ (distance from the star receiving the same flux density as Earth from the Sun). The three solid circles
mark three different photometric apertures used for the HOSTS survey. Simulations are shown, because this is typically
not visible in real data due to various noise terms that can only be removed by a statistical analysis.
The survey observations have now been completed, but a fraction of the newest observations are still being
analyzed. We discuss the latest results of our ongoing work and future prospects for the characterization of
exozodiacal dust with the LBTI.
2. OBSERVATIONS AND DATA ANALYSIS
A detailed description of our observations and data reduction can be found in [3]. We present here only a brief
summary.
Target stars have been selected for observations in real time from the full HOSTS target list presented by [19].
Our target stars are bright, nearby main sequence stars (Fν > 1 Jy in N band) without known close (< 1.5(cid:48)(cid:48))
binary companions. Spectral types range from A0 to K8. A list of observed stars with basic, relevant properties
is given in Table 1. Calibrators were selected following [2] using the catalogs of [20] and [21], supplemented by
stars from the JSDC catalog and the SearchCal tool (both [22]) where necessary.
The HOSTS survey was carried out with the LBTI on the Large Binocular Telescope (LBT) on Mt. Graham,
Arizona. The wavefronts from the two 8.4 m apertures -- separated by 14.4 m center to center -- are stabilized
using adaptive secondary mirrors. They are then brought to destructive interference in the LBTI and imaged on
the Nulling Optimized Mid-Infrared Camera (NOMIC, [39]). Tip-tilt and optical path delay (OPD) tracking are
performed using PhaseCam, a closed loop fringe tracker [40]. The adaptive optics system operates in the R to
I band range, PhaseCam in the K band, and nulling data are recorded in the N band. The interferometric inner
working angle is 70 mas. This is illustrated in Fig. 1. A nodding sequence is performed for sky subtraction and
photometric observations of each star are obtained for relative flux calibration. Science observations (SCI) are
paired with identical calibrator observations (CAL) to determine the nulling interferometric transfer function
(TF) of the instrument.
HD
Name
# SCI a
Sp. Type
V
K
number
Sensitivity driven sample:
(mag)
(mag)
Table 1. Observed stars included in this work.
N(cid:48)
(Jy)
d
(pc)
EEID fIR/nIR
(mas)
excess
Excess
references
β Eri
ζ Lep
33111
*38678
*81937
95418
97603
103287
106591
108767
128167
129502
*172167
187642
203280
Sun like stars sample:
23 UMa
β UMa
δ Leo
γ UMa
δ UMa
δ Crv
σ Boo
µ Vir
α Lyr
α Aql
α Cep
2
1
5
4
4
4
4
2
3
3
4
2
1
GJ 105 A
µ And
107 Psc
τ Cet
*9826
10476
*10700
16160
*22049
*30652
34411
48737
*78154
88230
89449
Eri
1 Ori
λ Aur
ξ Gem
13 UMa
GJ 380
40 Leo
β Vir
τ Boo
θ Boo
λ Ser
χ Her
γ Ser
2
3
2
1
4
4
2
3
3
2
2
4
*102870
4
*120136
3
126660
2
141004
3
142373
4
142860
5
*173667
2
185144
3
215648
2
*222368
Commissioning targets:
2
102647
109085
3
110 Her
σ Dra
ξ Peg A
ι Psc
β Leo
η Crv
A3 IV
A2 IV -- V
F0 IV
A1 IV
A5 IV
A0 V
A2 V
A0 IV
F4 V
F2 V
A0V
A7 V
A8 V
F9 V
K1 V
G8 V
K3 V
K2 V
F6 V
G1 V
F5 IV-V
F7 V
K8 V
F6 IV-V
F9 V
F6 IV
F7 V
G0 IV-V
G0 V
F6 IV
F6 V
G9 V
F6 V
F7 V
A3 V
F2 V
2.782
3.536
3.644
2.341
2.549
2.418
3.295
2.953
4.467
3.865
0.074
0.866
2.456
4.093
5.235
3.489
5.815
3.721
3.183
4.684
3.336
4.809
6.598
4.777
3.589
4.480
4.040
4.413
4.605
3.828
4.202
4.664
4.203
4.126
2.121
4.302
2.38
3.31
2.73
2.38
2.26
2.43
3.10
3.05
3.47
2.89
0.01
0.22
1.85
2.85
3.29
1.68
3.45
1.66
2.08
3.27
2.13
3.53
3.21
3.65
2.32
3.36
2.81
2.98
3.12
2.63
3.03
2.83
2.90
2.83
1.92
3.54
3.7
2.1
2.6
4.2
3.9
3.7
2.0
2.3
1.4
2.6
38.6
21.6
7.0
2.4
2.0
5.4
1.5
7.4
4.8
1.8
4.3
1.2
1.9
1.1
4.3
1.7
3.1
2.4
2.0
2.9
2.2
2.7
2.2
2.4
6.9
1.8
27.4
21.6
23.8
24.5
17.9
25.5
24.7
26.6
15.8
18.3
7.68
5.13
15.0
13.5
7.53
3.65
7.18
3.22
8.07
12.6
18.0
20.4
4.87
21.4
10.9
15.6
14.5
12.1
15.9
11.3
19.2
5.76
16.3
13.7
11.0
18.3
248
176
168
316
278
308
199
251
117
151
916
570
294
136
90
182
73
172
205
105
196
99
65
98
173
114
147
121
111
151
131
113
132
137
336
125
N/N
Y/N
N/ --
Y/N
N/N
N/ --
N/N
N/Y
Y/N
N/N
Y/Y
N/Y
N/Y
N/N
N/N
Y/Y
N/ --
Y/N
N/N
N/ --
-- /N
N/ --
N/ --
N/ --
N/N
N/N
N/ --
N/N
N/N
N/N
Y/Y
N/N
N/N
N/ --
Y/Y
Y/N
[6, 23, 24]
[6, 23]
[25]
[5, 26]
[5, 23, 24]
[23, 24, 26]
[5, 23, 24]
[6, 23, 24]
[5, 23]
[23, 24]
[4, 26]
[5, 23, 24, 27]
[5, 23, 24, 27]
[5, 25, 28]
[5, 23, 29, 30]
[5, 31]
[23, 29, 30]
[5, 32]
[5, 23, 29, 30]
[30, 33]
[5]
[23]
[28]
[23, 25]
[5, 30]
[6, 30, 33]
[23, 29, 30]
[5, 23, 29, 34]
[5, 23, 25, 29]
[5, 23, 29, 33]
[5, 28]
[5, 30, 33]
[23, 25, 29]
[23, 25, 29]
[5, 26]
[5, 35]
The full list of stars observed by the HOSTS survey and all null measurements and derived zodi levels will be
published in a dedicated paper (Ertel et al., in prep.). Stars with new observations since [3] are marked with an
asterisk. EEID is the Earth Equivalent Insolation Distance, the distance at which a body receives the same flux
density from its host star as Earth from the Sun. Magnitudes are given in the Vega system.
a Number of calibrated science pointings obtained. References are: Spectral type: SIMBAD; V magnitude:
[36]; K magnitude:
[37] and the Lausanne photometric data base (http://obswww.unige.ch/gcpd/gcpd.html);
N band flux and EEID: [19]; Distance: [38]
Figure 2. Distribution of excess significance Nas/σN (left) and uncertainties σN (right). The two vertical, dashed lines
in the excess significance distribution plots mark the ±3 σ boundaries based on our uncertainty estimates. The standard
deviation of the distribution is computed from non-detections only (−3 < Nas/σN < 3). The dotted line represents a
Gaussian with a standard deviation of one (Normal distribution) scaled to the peak of the histogram and is used only to
guide the eye.
After a basic reduction of each frame (nod subtraction, bad pixel correction), photometry is performed on
each single frame. The raw null depth and its uncertainty are determined using the null self calibration (NSC,
[41, 42, 43, 44]). Null measurements are converted to dust levels using a power-law radial dust distribution model
analogous to the Solar system's zodiacal dust [45, 46]. The vertical optical depth of the dust at a separation of
1 AU from the Sun of 7.12 × 10−8 defines our unit of 1 zodi.
The beam combination in the N band is achieved in the pupil plane before the light is re-imaged in the focal
plane. A consequence is that the light from a point source has the same phase over the entire detector, but
the phase is different at different sky angles. This results in a spatial filtering in form of a transmission pattern
(stripes of high and low transmission vertical to the baseline orientation) being applied to the flux distribution
on sky (Fig. 1).
3.1 Excess significance and detection threshold
3. RESULTS
Fig. 2 shows the distributions of the excess significance Nas/σN and of the uncertainties. The distributions are
in general well behaved with Nas/σN following a Normal distribution for −3 < Nas/σN < 3 with the addition
of several detections at Nas/σN > 3. We apply a 3 σ threshold to identify significant excesses. We find ten
significant excess detections.
3.2 Updates on specific targets
Seven of our ten detections have already been discussed in detail by [3]. Among those, no new data are included
in this work for β Leo, η Crv, β UMa, δ UMa, and θ Boo. New data have been taken and the results are included
here for Eri and 110 Her. Both detections were originally only achieved with a very large aperture size and
at relatively low significance. They thus required follow-up observations. For Eri we confirm our detection
with new, high accuracy data and measure a zodi level of 368±68 zodis. A detection is also achieved in a
smaller aperture (reduced background and read noise). The detection around 110 Her is also confirmed with
a new measurement of 343±57 zodis. This detection is, however, still only achieved in a large aperture, with
strong limits for smaller apertures, suggesting that the dust may be concentrated at a relatively large separation
from the star. As discussed already by [3] the low significance detection of far-infrared excess around this star
is controversial, so that it is not clear whether to include it in the group of stars with or without previously
detected cold dust.
−505101520Excess significance Nas/σN024681012NumberSigma f the distributi n: 1.19Excess distributi n0.00.20.40.60.81.0Abs lute uncertainty σN [%]024681012NumberMedian uncertainty: 0.09%Uncertainty distributionFigure 3. A HOSTS observing sequence (CAL -- SCI -- SCI -- CAL) on Vega showing the faint (4.5σ excess). The blue points
are calibrator observations, the red points are science observations. Measurements from each single nod are displayed.
The small vertical offset between science and calibrator measurements is the null excess. The first calibrator shown is
an unrelated observation of a star at a different sky location. The data illustrate the high instrument stability both over
time and over pointing direction.
In the new data obtained since summer 2017 (not included in [3]), we find three new detections: The Sun-like
star 13 UMa shows a very strong excess of 550±66 zodis. After δ UMa and θ Boo, this is the third star in our
survey without previously known cold excess, but with an LBTI detection (potentially the fourth one when
including 110 Her, see above). This further supports our conclusion in [3] that focussing on stars without known
cold dust for exo-Earth imaging does not guarantee a low HZ dust level. We further confirm the detection of HZ
dust around ζ Lep by [2] and measure a zodi level of 739±46 zodis.
Most interesting is the fact that new data obtained show a detection of 38±8 zodis around Vega. Our new
observations achieved a better sensitivity and the detection is consistent with the previous upper limit presented
in [3]. The excess is the strongest in a large aperture and not detected in a very small one, which means we are
not seeing the hot dust in this system previously reported by [4], but rather dust further out, near the HZ. An
observing sequence on Vega is shown in Fig. 3.
Finally, our upper limit on τ Cet of 74 zodis (3σ) is noteworthy. This star hosts a known debris disk (e.g.,
[47]), hot dust [5], and four claimed super-Earth mass planets (τ Cet e, f, g, h, [48]) with τ Cet f being in the
HZ, but also at sufficient angular separation that it might be detectable with a coronagraph on WFIRST. Our
upper limit shows that despite the high dust content in the system at large and very small separations, the HZ
is relatively dust poor.
3.3 Statistical results
In this section, we provide an update of the statistical analysis from [3] considering our additional data.
We divide our sample into early type stars and Sun-like stars, and into stars with previously known cold dust
('dusty stars') and without ('clean stars'). As in [3] we exclude η Crv and β Leo from the statistical analysis
as biased targets. Detection rates for these subsamples are listed in Table 2 and shown in Fig. 4. The results
of Fisher's exact test to check if differences in the detection rates are significant are shown in Table 3. We can
rule out with high confidence (probability 0.3%) that the occurrence rate of HZ dust is the same among clean
and dusty stars. Note that this result is affected by small number statistics of detections often at the boarder of
significance, both in our mid-infrared nulling data and in the far-infrared. For example, moving 110 Her into the
category of stars without a significant cold dust detection, this result moves to a considerably lower significance
probability changes to 0.017. The probability for having the same occurrence rate for stars with and without
known cold dust among Sun-like stars only is nominally 0.08, and becomes 0.32 when considering 110 Her as a
star without cold dust. We can thus not claim a significant correlation between cold and warm dust for Sun-like
stars and the correlation for all stars needs to be treated with care. The occurrence rates for Sun-like and
Figure 4. Occurrence rates of exozodiacal dust inferred from the observations considered in this work for early-type stars,
Sun-like stars, stars with previously known cold dust, and stars without detected cold dust.
Table 2. Subsamples, excess detections, and occurrence rates
Early
type
Sun-
like
All
Cold dust
3 of 4
75+10−27%
2 of 3
67+15−28%
5 of 7
71+11−20%
Clean
1 of 9
11+18−4 %
2 of 18
11+12−4 %
3 of 27
11+9−4%
All
4 of 13
31+15−9 %
4 of 21
19+11−6 %
8 of 34
23+9−6%
Table 3. Probability that two samples are drawn from the same distribution
Samples 1
All early type
All dusty
Clean early type
Clean Sun-like
Clean early type
Dusty early type
Sample 2
All Sun-like
All clean
Dusty early type
Dusty Sun-like
Clean Sun-like
Dusty Sun-like
Probability
0.24
0.003
0.05
0.08
0.47
0.57
early-type stars are the same within the binomial uncertainties. This is surprising given the typically ∼4 times
lower sensitivity to HZ dust surface density around Sun-like compared to early-type stars [19, 46]).
Next, we repeat the statistical analysis presented by [3] to determine the typical HZ dust levels around our
target stars, including the newest data available. We fit a lognormal exozodi luminosity function with parameters
µ and ς to our measurements using a maximum likelihood estimate and performing a Bayesian analysis. We
apply a 1/m prior, equivalent to assuming a flat prior in µ, marginalize the likelihood distribution over ς, and
compute the posterior cumulative probability distribution function (CPDF) of m = exp (µ). For Sun-like stars
without previously detected cold dust, we currently find a 95% confidence upper limit on the median zodi level
of 16 zodis. These stars are the highest priority targets for future exo-Earth imaging missions.
4. FUTURE PROSPECTS
4.1 Extending HOSTS and target vetting for future exo-Earth imaging
This summer 2018, the LBTI will undergo an adaptive optics upgrade designed to improve its wavefront stability
and performance at low target elevation. In addition, we are improving our application of the NSC to improve
background subtraction (currently a dominant source of uncertainty) with minimal change to the data acquisition
Figure 5. Simulation of observed null leak vs. hour angle of observation for two disks with different position angles and
inclinations. The disk model, the transmission pattern, and the transmitted signal are shown for three representative cases.
Over plotted on the i = 70 deg curve is a planned observing sequence and uncertainties for the proposed observations.
The high cadence, gray bars show the single nods and relative errors (relevant to detect variations over HA), while the
black points and uncertainties show the measurement when combining 6 nods including absolute calibrations (relevant
when combining observations from different nights). The large gaps are due to calibrator observations. Our detection
around β UMa is used as a template.
strategy. With the improved instrument performance and sensitivity, new survey observations will provide even
stronger statistical results.
Furthermore, our detections around θ Boo and 13 UMa (both Sun-like stars without known cold dust) show
that selecting such stars for exo-Earth imaging does not guarantee low HZ dust levels. One option to improve
over this selection strategy is to perform target vetting with the LBTI and a suitable future instrument in the
South [49, 50]. To illustrate the improvement possible with such a strategy, we produce from our sample of
Sun-like stars without cold dust an LBTI vetted sample by excluding the HOSTS detections. For such a sample
we find an upper limit on the median zodi level of 7.5 zodis, an improvement by a factor of two. This translates
in a 20% increase in yield for an exo-Earth imaging mission of a given size [15].
4.2 Characterization of detected systems using nulling interferometry
The performance of LBTI for the HOSTS survey has reached a level that makes nulling infrared interferometry
a powerful tool to study warm dust systems around nearby main sequence stars. With a detection rate of the
HOSTS survey of ∼25% and a sensitivity of only a few zodis for our most favorable stars, we can now study
systems with typical HZ dust levels for the first time. Our high signal-to-noise detections allow for detailed
follow-up observations.
The alt-azimuth mount of the LBTI allows us to measure the excess at different position angles of the single
interferometric baseline projected on sky when observing at different hour angles. A different disk size along
different position angles (an inclined disk) or disk structure cause variation of the null excess with baseline
orientation. This allows us to measure the inclination and position angle of the disk, and to search for structures
due to planet-disk interaction. The latter will rotate with the orbital period of the responsible planet, which is
of the order of a few years for planets close to the HZ. Thus, monitoring over a few years can be used to to better
distinguish between different scenarios for the hour angle variation and to constrain the orbit of the responsible
planet. In addition, observations at different wavelengths across the N band provide different spatial filtering
by the transmission pattern, thus further constraining the geometry of the emission. Multi-wavelength data also
allow us to constrain the spectral shape of the excess emission.
These information allow us to address the most pressing questions on exozodiacal dust after the HOSTS
survey. Determining the geometry and radial dust distribution is critical for an accurate determination of the
zodi level from our nulling data. Information on the radial dust distribution for a range of systems with different
zodi levels and stellar spectral types will also allow us to make better estimates of the probable levels of exozodi
in systems where it has not been detected. This will allow us to derive more reliable upper limits and thus more
reliable statistics from the HOSTS survey. The radial slope of the dust distribution can inform us about the dust
origin: Poynting-Robertson drag from an outer disk would result in a flat or inwardly decreasing surface density
distribution [11], while a cometary origin would produce an inwardly increasing one [51, 52]. The detection of
potential disk structures and radial discontinuities of the surface density distribution allows us to put constraints
on the responsible planets (orbit, mass, [13, 14], Bonsor et al., in prep.). This provides present-day insight into
the architecture and dynamics of nearby planetary systems in and near their HZs. Information on the spectral
slope of the dust emission allows us to constrain the dust grain size [53] and thus its origin (Poynting-Robertson
drag, local collisional production, or cometary origin). Moreover, a better understanding of the dust grain size
allows us to better predict its scattered light brightness [54]. This is critical to better understand the implications
of our HOSTS results (obtained in the mid-infrared) for exo-Earth imaging (to be achieved in the visible).
We demonstrate the feasibility of the proposed observations using simulations of the expected signal for
several scenarios, using our detection around β UMa as a template. Fig. 5 shows the null excess vs. hour angle of
observations for two disks at different inclinations and position angles. It can be seen that the null excess varies
by ∼20% and ∼35% for the two examples due to the disk inclination. Fig. 6 shows the structure produced by
a 20 M⊕ planet at 1 AU from a Sun-like star in dust spiraling inward from an asteroid belt at 2.25 AU due to
Poynting-Robertson drag [14] and the expected null variation vs. position angle of LBTI's transmission pattern.
Again, a variation of ∼20% can be expected. LBTI's current accuracy to null excess is ∼ 6 × 10−4 (1 σ) for a
calibrated observing sequence. Thus, the expected signals can be detected with two observations in the extrema
of the null variation if the mean excess is ∼1%. They should thus be well detectable in our systems with the
strongest excesses (β Leo, β UMa, η Crv, ζ Lep, 13 UMa). In addition, strong structures on top of the measured
excesses may be visible for our other detections including Eri and Vega. Because the hour angles of the extrema
of the null excess are unknown a priori, observations need to be carried out at a range of hour angles. This will
naturally produce more data than observations in the extrema only and thus yield a higher sensitivity to the
observed features. A realistic observing cadence and hour angle coverage of observations optimized to detect
variations of the null excess with hour angle is shown in Fig. 5.
Simulations at two wavelengths (e.g., 7.9 µm and 11.9 µm, filter width ∼10%) of a face-on disk with two
different surface density slopes (flat and outward decreasing with 1/r) are shown in Fig. 7. From these data
one can measure the null excess with different photometric aperture sizes to reconstruct its angular size which
is at least marginally resolved by the single aperture PSF. Moreover, one can measure the null excess ratio
between two wavelengths, which is N11.9/N7.9 = 1.4 and N11.9/N7.9 = 0.9 for the flat and 1/r surface density
cases, respectively. This difference is also detectable for null excesses around 1% with single observations in each
wavelength, and for lower excesses if more data are obtained.
Detailed model fitting, combining the wealth of data provided by the observations described above can also
break degeneracies in single data sets [46, 53, 54, 55]. Such degeneracies are for example present between disk
inclination and structure as a source of hour angle variation of the measured excess or between dust grain size
and radial distribution in the wavelength dependence of the measured null excesses. However, the first can be
broken when measuring the variation at two wavelengths (the inclination effect is related to the inner working
angle, the effect of a clump typically not). The latter can be broken by measuring the radial distribution also
with different photometric aperture sizes.
Figure 6. Simulation of observed null leak vs. hour angle (right) for an observation of a disk with resonant structures
due to planet-disk-interaction (left, [14]). The dust surface density is shown in the left image, where red corresponds to
high and blue to low surface density. Simulations have been performed for a 20 M⊕ planet at a separation of 1 au from a
Sun-like star (distance 10 pc).
Figure 7. Transmission of disk flux at 7.9 µm (top row ) and 11.9 µm (bottom row ) for two disk models with different radial
surface density (left three columns: flat, right three: 1/r). The disk model image (in logarithmic scale), transmission
pattern, and transmitted flux are shown. All disk model images are shown in the same color scale. Transmitted flux is
shown in the same color scale for the same wavelength but the peak brightness is twice as high at 7.9 µm than at 11.9 µm.
The flux ratio of the null excess at 11.9 µm and 7.9 µm is N11.9/N7.9 = 1.4 for the flat surface density and N11.9/N7.9 = 0.9
for the 1/r case. The white circles illustrate the FWHM of the single telescope PSF. It can be seen that the transmitted
flux is resolved. Our detection of β UMa is used as a template.
4.3 Characterization of detected systems using Fizeau spectro-interferometry
Another possibility the LBTI offers for the characterization of exozodiacal dust is spectro-interferometry in the
K to N band in constructive Fizeau mode. The LBTI is currently the only interferometer operating in this
wavelength range and using a fringe tracker. Thus, it should in principle be possible to obtain spectrally resolved
data with an accuracy sufficient to detect excesses at the ∼1% level. Such data could for example be used to
study the connection between HZ dust and hot dust closer in by tracing the warm dust emission toward shorter
wavelengths of the hot emission toward longer wavelengths. It should also be possible to detect the 3 µm and
10 µm silicate features of the dust to further study its composition. A description of LBTI's Fizeau mode and
our efforts in implementing routine, accurate, and efficient observations is presented by [56].
We have conducted the most sensitive survey for HZ dust around nearby stars. We put strong constraints on the
typical HZ dust levels and show that these are low enough for stars without known cold dust to allow exo-Earth
5. CONCLUSIONS
imaging. However, our detections around these stars demonstrate that such a target selection strategy does not
guarantee low HZ dust levels for all stars. This limits our constraint on the median dust level to 16 zodis, while
for a sample of LBTI vetted stars the median level would be by a factor of two better. This transforms into a
20% increase of the yield of an exo-Earth imaging mission of a given size. We have also outlined the science case
for a future characterization of the HOSTS detected exozodis with the LBTI and demonstrated the feasibility
of the proposed observations through modeling of the expected signals. These observations represent a critical
next step in the understanding of the origin and properties of exozodiacal dust and provide present-day insights
into the architecture and dynamics of nearby planetary systems near their HZs.
ACKNOWLEDGMENTS
The Large Binocular Telescope Interferometer is funded by the National Aeronautics and Space Administration
as part of its Exoplanet Exploration Program. The LBT is an international collaboration among institutions in
the United States, Italy, and Germany. LBT Corporation partners are: The University of Arizona on behalf of
the Arizona university system; Instituto Nazionale di Astrofisica, Italy; LBT Beteiligungsgesellschaft, Germany,
representing the Max-Planck Society, the Astrophysical Institute Potsdam, and Heidelberg University; The Ohio
State University, and The Research Corporation, on behalf of The University of Notre Dame, University of
Minnesota and University of Virginia. This research has made extensive use of the SIMBAD database [57] and
the VizieR catalogue access tool [58], both operated at CDS, Strasbourg, France, of Python, including the NumPy,
SciPy, Matplotlib [59], and Astorpy [60] libraries, and of NASA's Astrophysics Data System Bibliographic
Services. GMK is supported by the Royal Society as a Royal Society University Research Fellow. AS is partially
supported by funding from the Center for Exoplanets and Habitable Worlds. The Center for Exoplanets and
Habitable Worlds is supported by the Pennsylvania State University, the Eberly College of Science, and the
Pennsylvania Space Grant Consortium. JMS is supported by NASA through Hubble Fellowship grant HST-
HF2-51398.001-A awarded by the Space Telescope Science Institute, which is operated by the Association of
Universities for Research in Astronomy, Inc., for NASA, under contract NAS5-26555.
References
[1] Millan-Gabet, R., Serabyn, E., Mennesson, B., Stark, C. C., Ragland, S., Hrynevych, M., Woillez, J.,
Stapelfeldt, K., Bryden, G., Colavita, M. M., and Booth, A. J., "Exozodiacal Dust Levels for Nearby
Main-sequence Stars: A Survey with the Keck Interferometer Nuller," ApJ 734, 67 (June 2011).
[2] Mennesson, B., Millan-Gabet, R., Serabyn, E., Colavita, M. M., Absil, O., Bryden, G., Wyatt, M., Danchi,
W., Defr`ere, D., Dor´e, O., Hinz, P., Kuchner, M., Ragland, S., Scott, N., Stapelfeldt, K., Traub, W., and
Woillez, J., "Constraining the Exozodiacal Luminosity Function of Main-sequence Stars: Complete Results
from the Keck Nuller Mid-infrared Surveys," ApJ 797, 119 (Dec. 2014).
[3] Ertel, S., Defr`ere, D., Hinz, P., Mennesson, B., Kennedy, G. M., Danchi, W. C., Gelino, C., Hill, J. M.,
Hoffmann, W. F., Rieke, G., Shannon, A., Spalding, E., Stone, J. M., Vaz, A., Weinberger, A. J., Willems,
P., Absil, O., Arbo, P., Bailey, V. P., Beichman, C., Bryden, G., Downey, E. C., Durney, O., Esposito,
S., Gaspar, A., Grenz, P., Haniff, C. A., Leisenring, J. M., Marion, L., McMahon, T. J., Millan-Gabet, R.,
Montoya, M., Morzinski, K. M., Pinna, E., Power, J., Puglisi, A., Roberge, A., Serabyn, E., Skemer, A. J.,
Stapelfeldt, K., Su, K. Y. L., Vaitheeswaran, V., and Wyatt, M. C., "The HOSTS Survey -- Exozodiacal
Dust Measurements for 30 Stars," AJ 155, 194 (May 2018).
[4] Absil, O., di Folco, E., M´erand, A., Augereau, J.-C., Coud´e du Foresto, V., Aufdenberg, J. P., Kervella, P.,
Ridgway, S. T., Berger, D. H., ten Brummelaar, T. A., Sturmann, J., Sturmann, L., Turner, N. H., and
McAlister, H. A., "Circumstellar material in the ¡ASTROBJ¿Vega¡/ASTROBJ¿ inner system revealed by
CHARA/FLUOR," A&A 452, 237 -- 244 (June 2006).
[5] Absil, O., Defr`ere, D., Coud´e du Foresto, V., Di Folco, E., M´erand, A., Augereau, J.-C., Ertel, S., Hanot, C.,
Kervella, P., Mollier, B., Scott, N., Che, X., Monnier, J. D., Thureau, N., Tuthill, P. G., ten Brummelaar,
T. A., McAlister, H. A., Sturmann, J., Sturmann, L., and Turner, N., "A near-infrared interferometric survey
of debris-disc stars. III. First statistics based on 42 stars observed with CHARA/FLUOR," A&A 555, A104
(July 2013).
[6] Ertel, S., Absil, O., Defr`ere, D., Le Bouquin, J.-B., Augereau, J.-C., Marion, L., Blind, N., Bonsor, A.,
Bryden, G., Lebreton, J., and Milli, J., "A near-infrared interferometric survey of debris-disk stars. IV. An
unbiased sample of 92 southern stars observed in H band with VLTI/PIONIER," A&A 570, A128 (Oct.
2014).
[7] Ertel, S., Defr`ere, D., Absil, O., Le Bouquin, J.-B., Augereau, J.-C., Berger, J.-P., Blind, N., Bonsor, A.,
Lagrange, A.-M., Lebreton, J., Marion, L., Milli, J., and Olofsson, J., "A near-infrared interferometric
survey of debris-disc stars. V. PIONIER search for variability," A&A 595, A44 (Oct. 2016).
[8] Kral, Q., Krivov, A. V., Defr`ere, D., van Lieshout, R., Bonsor, A., Augereau, J.-C., Th´ebault, P., Ertel,
S., Lebreton, J., and Absil, O., "Exozodiacal clouds: hot and warm dust around main sequence stars," The
Astronomical Review 13, 69 -- 111 (Apr. 2017).
[9] Faramaz, V., Ertel, S., Booth, M., Cuadra, J., and Simmonds, C., "Inner mean-motion resonances with
eccentric planets: a possible origin for exozodiacal dust clouds," MNRAS 465, 2352 -- 2365 (Feb. 2017).
[10] Backman, D. E. and Paresce, F., "Main-sequence stars with circumstellar solid material - The VEGA
phenomenon," in [Protostars and Planets III ], E. H. Levy & J. I. Lunine, ed., 1253 -- 1304 (1993).
[11] Kennedy, G. M. and Piette, A., "Warm exo-Zodi from cool exo-Kuiper belts: the significance of P-R drag
and the inference of intervening planets," MNRAS 449, 2304 -- 2311 (May 2015).
[12] Reidemeister, M., Krivov, A. V., Stark, C. C., Augereau, J., Lohne, T., and Muller, S., "The cold origin of
the warm dust around Eridani," A&A 527, A57+ (Mar. 2011).
[13] Ertel, S., Wolf, S., and Rodmann, J., "Observing planet-disk interaction in debris disks," A&A 544, A61
(Aug. 2012).
[14] Shannon, A., Mustill, A. J., and Wyatt, M., "Capture and evolution of dust in planetary mean-motion
resonances: a fast, semi-analytic method for generating resonantly trapped disc images," MNRAS 448,
684 -- 702 (Mar. 2015).
[15] Stark, C. C., Roberge, A., Mandell, A., Clampin, M., Domagal-Goldman, S. D., McElwain, M. W., and
Stapelfeldt, K. R., "Lower Limits on Aperture Size for an ExoEarth Detecting Coronagraphic Mission,"
ApJ 808, 149 (Aug. 2015).
[16] Stark, C. C., Shaklan, S., Lisman, D., Cady, E., Savransky, D., Roberge, A., and Mandell, A. M., "Max-
imized exoEarth candidate yields for starshades," Journal of Astronomical Telescopes, Instruments, and
Systems 2, 041204 (Oct. 2016).
[17] Defr`ere, D., Absil, O., den Hartog, R., Hanot, C., and Stark, C., "Nulling interferometry:
impact of
exozodiacal clouds on the performance of future life-finding space missions," A&A 509, A9 (Jan. 2010).
[18] Hinz, P. M., Defr`ere, D., Skemer, A., Bailey, V., Stone, J., Spalding, E., Vaz, A., Pinna, E., Puglisi, A.,
Esposito, S., Montoya, M., Downey, E., Leisenring, J., Durney, O., Hoffmann, W., Hill, J., Millan-Gabet,
R., Mennesson, B., Danchi, W., Morzinski, K., Grenz, P., Skrutskie, M., and Ertel, S., "Overview of LBTI:
a multipurpose facility for high spatial resolution observations," in [Optical and Infrared Interferometry and
Imaging V], Proc. SPIE 9907, 990704 (Aug. 2016).
[19] Weinberger, A. J., Bryden, G., Kennedy, G. M., Roberge, A., Defr`ere, D., Hinz, P. M., Millan-Gabet, R.,
Rieke, G., Bailey, V. P., Danchi, W. C., Haniff, C., Mennesson, B., Serabyn, E., Skemer, A. J., Stapelfeldt,
K. R., and Wyatt, M. C., "Target Selection for the LBTI Exozodi Key Science Program," ApJS 216, 24
(Feb. 2015).
[20] Bord´e, P., Coud´e du Foresto, V., Chagnon, G., and Perrin, G., "A catalogue of calibrator stars for long
baseline stellar interferometry," A&A 393, 183 -- 193 (Oct. 2002).
[21] M´erand, A., Bord´e, P., and Coud´e du Foresto, V., "A catalog of bright calibrator stars for 200-m baseline
near-infrared stellar interferometry," A&A 433, 1155 -- 1162 (Apr. 2005).
[22] Chelli, A., Duvert, G., Bourg`es, L., Mella, G., Lafrasse, S., Bonneau, D., and Chesneau, O., "Pseudomag-
nitudes and differential surface brightness: Application to the apparent diameter of stars," A&A 589, A112
(May 2016).
[23] G´asp´ar, A., Rieke, G. H., and Balog, Z., "The Collisional Evolution of Debris Disks," ApJ 768, 25 (May
2013).
[24] Thureau, N. D., Greaves, J. S., Matthews, B. C., Kennedy, G., Phillips, N., Booth, M., Duchene, G., Horner,
J., Rodriguez, D. R., Sibthorpe, B., and Wyatt, M. C., "An unbiased study of debris discs around A-type
stars with Herschel," MNRAS 445, 2558 -- 2573 (Dec. 2014).
[25] Beichman, C. A., Bryden, G., Stapelfeldt, K. R., Gautier, T. N., Grogan, K., Shao, M., Velusamy, T.,
Lawler, S. M., Blaylock, M., Rieke, G. H., Lunine, J. I., Fischer, D. A., Marcy, G. W., Greaves, J. S.,
Wyatt, M. C., Holland, W. S., and Dent, W. R. F., "New Debris Disks around Nearby Main-Sequence
Stars: Impact on the Direct Detection of Planets," ApJ 652, 1674 -- 1693 (Dec. 2006).
[26] Su, K. Y. L., Rieke, G. H., Stansberry, J. A., Bryden, G., Stapelfeldt, K. R., Trilling, D. E., Muzerolle, J.,
Beichman, C. A., Moro-Martin, A., Hines, D. C., and Werner, M. W., "Debris Disk Evolution around A
Stars," ApJ 653, 675 -- 689 (Dec. 2006).
[27] Rieke, G. H., Su, K. Y. L., Stansberry, J. A., Trilling, D., Bryden, G., Muzerolle, J., White, B., Gorlova, N.,
Young, E. T., Beichman, C. A., Stapelfeldt, K. R., and Hines, D. C., "Decay of Planetary Debris Disks,"
ApJ 620, 1010 -- 1026 (Feb. 2005).
[28] Eiroa, C., Marshall, J. P., Mora, A., Montesinos, B., Absil, O., Augereau, J. C., Bayo, A., Bryden, G.,
Danchi, W., del Burgo, C., Ertel, S., Fridlund, M., Heras, A. M., Krivov, A. V., Launhardt, R., Liseau,
R., Lohne, T., Maldonado, J., Pilbratt, G. L., Roberge, A., Rodmann, J., Sanz-Forcada, J., Solano, E.,
Stapelfeldt, K., Th´ebault, P., Wolf, S., Ardila, D., Ar´evalo, M., Beichmann, C., Faramaz, V., Gonz´alez-
Garc´ıa, B. M., Guti´errez, R., Lebreton, J., Mart´ınez-Arn´aiz, R., Meeus, G., Montes, D., Olofsson, G.,
Su, K. Y. L., White, G. J., Barrado, D., Fukagawa, M., Grun, E., Kamp, I., Lorente, R., Morbidelli, A.,
Muller, S., Mutschke, H., Nakagawa, T., Ribas, I., and Walker, H., "DUst around NEarby Stars. The survey
observational results," A&A 555, A11 (July 2013).
[29] Montesinos, B., Eiroa, C., Krivov, A. V., Marshall, J. P., Pilbratt, G. L., Liseau, R., Mora, A., Maldonado,
J., Wolf, S., Ertel, S., Bayo, A., Augereau, J.-C., Heras, A. M., Fridlund, M., Danchi, W. C., Solano, E.,
Kirchschlager, F., del Burgo, C., and Montes, D., "Incidence of debris discs around FGK stars in the solar
neighbourhood," A&A 593, A51 (Sept. 2016).
[30] Trilling, D. E., Bryden, G., Beichman, C. A., Rieke, G. H., Su, K. Y. L., Stansberry, J. A., Blaylock,
M., Stapelfeldt, K. R., Beeman, J. W., and Haller, E. E., "Debris Disks around Sun-like Stars," ApJ 674,
1086 -- 1105 (Feb. 2008).
[31] Greaves, J. S., Wyatt, M. C., Holland, W. S., and Dent, W. R. F., "The debris disc around τ Ceti: a massive
analogue to the Kuiper Belt," MNRAS 351, L54 -- L58 (July 2004).
[32] Aumann, H. H., "IRAS observations of matter around nearby stars," PASP 97, 885 -- 891 (Oct. 1985).
[33] Lawler, S. M., Beichman, C. A., Bryden, G., Ciardi, D. R., Tanner, A. M., Su, K. Y. L., Stapelfeldt, K. R.,
Lisse, C. M., and Harker, D. E., "Explorations Beyond the Snow Line: Spitzer/IRS Spectra of Debris Disks
Around Solar-type Stars," ApJ 705, 89 -- 111 (Nov. 2009).
[34] Koerner, D. W., Kim, S., Trilling, D. E., Larson, H., Cotera, A., Stapelfeldt, K. R., Wahhaj, Z., Fajardo-
Acosta, S., Padgett, D., and Backman, D., "New Debris Disk Candidates Around 49 Nearby Stars," ApJ 710,
L26 -- L29 (Feb. 2010).
[35] Aumann, H. H., "Spectral class distribution of circumstellar material in main-sequence stars," AJ 96,
1415 -- 1419 (Oct. 1988).
[36] Kharchenko, N. V., Scholz, R., Piskunov, A. E., Roeser, S., and Schilbach, E., "2nd Cat. of Radial Velocities
with Astrometric Data (Kharchenko+, 2007)," VizieR Online Data Catalog 3254, 0 -- + (June 2007).
[37] Gezari, D. Y., Schmitz, M., Pitts, P. S., and Mead, J. M., [Catalog of infrared observations, third edition ]
(June 1993).
[38] van Leeuwen, F., "Validation of the new Hipparcos reduction," A&A 474, 653 -- 664 (Nov. 2007).
[39] Hoffmann, W. F., Hinz, P. M., Defr`ere, D., Leisenring, J. M., Skemer, A. J., Arbo, P. A., Montoya, M., and
Mennesson, B., "Operation and performance of the mid-infrared camera, NOMIC, on the Large Binocular
Telescope," in [Ground-based and Airborne Instrumentation for Astronomy V], Proc. SPIE 9147, 91471O
(July 2014).
[40] Defr`ere, D., Hinz, P., Downey, E., Ashby, D., Bailey, V., Brusa, G., Christou, J., Danchi, W. C., Grenz,
P., Hill, J. M., Hoffmann, W. F., Leisenring, J., Lozi, J., McMahon, T., Mennesson, B., Millan-Gabet, R.,
Montoya, M., Powell, K., Skemer, A., Vaitheeswaran, V., Vaz, A., and Veillet, C., "Co-phasing the Large
Binocular Telescope: status and performance of LBTI/PHASECam," ArXiv e-prints (Jan. 2015).
[41] Mennesson, B., Serabyn, E., Hanot, C., Martin, S. R., Liewer, K., and Mawet, D., "New Constraints on
Companions and Dust within a Few AU of Vega," ApJ 736, 14 (July 2011).
[42] Hanot, C., Mennesson, B., Martin, S., Liewer, K., Loya, F., Mawet, D., Riaud, P., Absil, O., and Serabyn,
E., "Improving Interferometric Null Depth Measurements using Statistical Distributions: Theory and First
Results with the Palomar Fiber Nuller," ApJ 729, 110 (Mar. 2011).
[43] Defr`ere, D., Hinz, P. M., Mennesson, B., Hoffmann, W. F., Millan-Gabet, R., Skemer, A. J., Bailey, V.,
Danchi, W. C., Downey, E. C., Durney, O., Grenz, P., Hill, J. M., McMahon, T. J., Montoya, M., Spalding,
E., Vaz, A., Absil, O., Arbo, P., Bailey, H., Brusa, G., Bryden, G., Esposito, S., Gaspar, A., Haniff, C. A.,
Kennedy, G. M., Leisenring, J. M., Marion, L., Nowak, M., Pinna, E., Powell, K., Puglisi, A., Rieke, G.,
Roberge, A., Serabyn, E., Sosa, R., Stapeldfeldt, K., Su, K., Weinberger, A. J., and Wyatt, M. C., "Nulling
Data Reduction and On-sky Performance of the Large Binocular Telescope Interferometer," ApJ 824, 66
(June 2016).
[44] Mennesson, B., Defr`ere, D., Nowak, M., Hinz, P., Millan-Gabet, R., Absil, O., Bailey, V., Bryden, G.,
Danchi, W., Kennedy, G. M., Marion, L., Roberge, A., Serabyn, E., Skemer, A. J., Stapelfeldt, K., Wein-
berger, A. J., and Wyatt, M., "Making high-accuracy null depth measurements for the LBTI exozodi survey,"
in [Optical and Infrared Interferometry and Imaging V], Proc. SPIE 9907, 99070X (Aug. 2016).
[45] Kelsall, T., Weiland, J. L., Franz, B. A., Reach, W. T., Arendt, R. G., Dwek, E., Freudenreich, H. T.,
Hauser, M. G., Moseley, S. H., Odegard, N. P., Silverberg, R. F., and Wright, E. L., "The COBE Diffuse
Infrared Background Experiment Search for the Cosmic Infrared Background. II. Model of the Interplanetary
Dust Cloud," ApJ 508, 44 -- 73 (Nov. 1998).
[46] Kennedy, G. M., Wyatt, M. C., Bailey, V., Bryden, G., Danchi, W. C., Defr`ere, D., Haniff, C., Hinz, P. M.,
Lebreton, J., Mennesson, B., Millan-Gabet, R., Morales, F., Pani´c, O., Rieke, G. H., Roberge, A., Serabyn,
E., Shannon, A., Skemer, A. J., Stapelfeldt, K. R., Su, K. Y. L., and Weinberger, A. J., "Exo-zodi Modeling
for the Large Binocular Telescope Interferometer," ApJS 216, 23 (Feb. 2015).
[47] Lawler, S. M., Di Francesco, J., Kennedy, G. M., Sibthorpe, B., Booth, M., Vandenbussche, B., Matthews,
B. C., Holland, W. S., Greaves, J., Wilner, D. J., Tuomi, M., Blommaert, J. A. D. L., de Vries, B. L.,
Dominik, C., Fridlund, M., Gear, W., Heras, A. M., Ivison, R., and Olofsson, G., "The debris disc of solar
analogue τ Ceti: Herschel observations and dynamical simulations of the proposed multiplanet system,"
MNRAS 444, 2665 -- 2675 (Nov. 2014).
[48] Feng, F., Tuomi, M., Jones, H. R. A., Barnes, J., Anglada-Escud´e, G., Vogt, S. S., and Butler, R. P., "Color
Difference Makes a Difference: Four Planet Candidates around τ Ceti," AJ 154, 135 (Oct. 2017).
[49] Defr`ere, D., Absil, O., Berger, J.-P., Boulet, T., Danchi, W. C., Ertel, S., Gallenne, A., H´enault, F., Hinz,
P., Huby, E., Ireland, M., Kraus, S., Labadie, L., Le Bouquin, J.-B., Martin, G., Matter, A., M´erand, A.,
Mennesson, B., Minardi, S., Monnier, J., Norris, B., Orban de Xivry, G., Pedretti, E., Pott, J.-U., Reggiani,
M., Serabyn, E., Surdej, J., Tristram, K. R. W., and Woillez, J., "The path towards high-contrast imaging
with the VLTI: the Hi-5 project," ArXiv e-prints (Jan. 2018).
[50] Defr`ere, D., Absil, O., Berger, J.-P., Boulet, T., Danchi, W. C., Ertel, S., Gallenne, A., H´enault, F., Hinz,
P., Huby, E., Ireland, M., Kraus, S., Labadie, L., Le Bouquin, J.-B., Martin, G., Matter, A., M´erand, A.,
Mennesson, B., Minardi, S., Monnier, J., Norris, B., Orban de Xivry, G., Pedretti, E., Pott, J.-U., Reggiani,
M., Serabyn, E., Surdej, J., Tristram, K. R. W., and Woillez, J., "Hi-5: a potential high-contrast thermal
near-infrared imager for the VLTI," To appear in the same issue as the present paper. (2018).
[51] Nesvorn´y, D., Jenniskens, P., Levison, H. F., Bottke, W. F., Vokrouhlick´y, D., and Gounelle, M., "Cometary
Origin of the Zodiacal Cloud and Carbonaceous Micrometeorites. Implications for Hot Debris Disks,"
ApJ 713, 816 -- 836 (Apr. 2010).
[52] Marshall, J. P., Cotton, D. V., Bott, K., Ertel, S., Kennedy, G. M., Wyatt, M. C., del Burgo, C., Absil,
O., Bailey, J., and Kedziora-Chudczer, L., "Polarization Measurements of Hot Dust Stars and the Local
Interstellar Medium," ApJ 825, 124 (July 2016).
[53] Ertel, S., Wolf, S., Marshall, J. P., Eiroa, C., Augereau, J.-C., Krivov, A. V., Lohne, T., Absil, O., Ardila,
D., Ar´evalo, M., Bayo, A., Bryden, G., del Burgo, C., Greaves, J., Kennedy, G., Lebreton, J., Liseau,
R., Maldonado, J., Montesinos, B., Mora, A., Pilbratt, G. L., Sanz-Forcada, J., Stapelfeldt, K., and White,
G. J., "A peculiar class of debris disks from Herschel/DUNES. A steep fall off in the far infrared," A&A 541,
A148 (May 2012).
[54] Ertel, S., Wolf, S., Metchev, S., Schneider, G., Carpenter, J. M., Meyer, M. R., Hillenbrand, L. A., and Sil-
verstone, M. D., "Multi-wavelength modeling of the spatially resolved debris disk of HD 107146," A&A 533,
A132+ (Sept. 2011).
[55] Ertel, S., Marshall, J. P., Augereau, J.-C., Krivov, A. V., Lohne, T., Eiroa, C., Mora, A., del Burgo, C.,
Montesinos, B., Bryden, G., Danchi, W., Kirchschlager, F., Liseau, R., Maldonado, J., Pilbratt, G. L.,
Schuppler, C., Th´ebault, P., White, G. J., and Wolf, S., "Potential multi-component structure of the debris
disk around HIP 17439 revealed by Herschel/DUNES," A&A 561, A114 (Jan. 2014).
[56] Spalding, E., Hinz, P., Ertel, S., Maier, E., and Stone, J., "Towards controlled Fizeau observations with the
Large Binocular Telescope," To appear in the same issue as the present paper. (2018).
[57] Wenger, M., Ochsenbein, F., Egret, D., Dubois, P., Bonnarel, F., Borde, S., Genova, F., Jasniewicz, G.,
Laloe, S., Lesteven, S., and Monier, R., "The SIMBAD astronomical database. The CDS reference database
for astronomical objects," A&AS 143, 9 -- 22 (Apr. 2000).
[58] Ochsenbein, F., Bauer, P., and Marcout, J., "The VizieR database of astronomical catalogues," A&AS 143,
23 -- 32 (Apr. 2000).
[59] Hunter, J. D., "Matplotlib: A 2d graphics environment," Computing In Science & Engineering 9(3), 90 -- 95
(2007).
[60] Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., Greenfield, P., Droettboom, M., Bray, E., Aldcroft,
T., Davis, M., Ginsburg, A., Price-Whelan, A. M., Kerzendorf, W. E., Conley, A., Crighton, N., Barbary,
K., Muna, D., Ferguson, H., Grollier, F., Parikh, M. M., Nair, P. H., Unther, H. M., Deil, C., Woillez,
J., Conseil, S., Kramer, R., Turner, J. E. H., Singer, L., Fox, R., Weaver, B. A., Zabalza, V., Edwards,
Z. I., Azalee Bostroem, K., Burke, D. J., Casey, A. R., Crawford, S. M., Dencheva, N., Ely, J., Jenness, T.,
Labrie, K., Lim, P. L., Pierfederici, F., Pontzen, A., Ptak, A., Refsdal, B., Servillat, M., and Streicher, O.,
"Astropy: A community Python package for astronomy," A&A 558, A33 (Oct. 2013).
|
1602.07457 | 1 | 1602 | 2016-02-24T10:25:19 | Two mechanisms for dust gap opening in protoplanetary discs | [
"astro-ph.EP"
] | We identify two distinct physical mechanisms for dust gap opening by embedded planets in protoplanetary discs based on the symmetry of the drag-induced motion around the planet: I) A mechanism where low mass planets, that do not disturb the gas, open gaps in dust by tidal torques assisted by drag in the inner disc, but resisted by drag in the outer disc; and II) The usual, drag assisted, mechanism where higher mass planets create pressure maxima in the gas disc which the drag torque then acts to evacuate further in the dust. The first mechanism produces gaps in dust but not gas, while the second produces partial or total gas gaps which are deeper in the dust phase. Dust gaps do not necessarily indicate gas gaps. | astro-ph.EP | astro-ph | MNRAS 000, 1 -- 5 (2016)
Preprint 11 September 2018
Compiled using MNRAS LATEX style file v3.0
Two mechanisms for dust gap opening in protoplanetary discs
Giovanni Dipierro1 ⋆, Guillaume Laibe2, Daniel J. Price3 and Giuseppe Lodato1
1Dipartimento di Fisica, Universit`a Degli Studi di Milano, Via Celoria, 16, Milano, I-20133, Italy
2School of Physics and Astronomy, University of St. Andrews, North Haugh, St. Andrews, Fife KY16 9SS, UK
3Monash Centre for Astrophysics (MoCA) and School of Physics and Astronomy, Monash University, Clayton Vic 3800, Australia
6
1
0
2
b
e
F
4
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
7
5
4
7
0
.
2
0
6
1
:
v
i
X
r
a
ABSTRACT
We identify two distinct physical mechanisms for dust gap opening by embedded planets in
protoplanetary discs based on the symmetry of the drag-induced motion around the planet:
I) A mechanism where low mass planets, that do not disturb the gas, open gaps in dust by
tidal torques assisted by drag in the inner disc, but resisted by drag in the outer disc; and II)
The usual, drag assisted, mechanism where higher mass planets create pressure maxima in the
gas disc which the drag torque then acts to evacuate further in the dust. The first mechanism
produces gaps in dust but not gas, while the second produces partial or total gas gaps which
are deeper in the dust phase. Dust gaps do not necessarily indicate gas gaps.
Key words:
submillimetre: planetary systems
protoplanetary discs -- planet-disc interactions -- dust, extinction --
1 INTRODUCTION
spectacular
Recent
spatially resolved observations of gaps
and ring-like structures in nearby dusty protoplanetary discs
(ALMA Partnership et al. 2015; Nomura et al. 2015) have revived
interest in studying gap-opening mechanisms in discs and in ex-
tending the theoretical and numerical investigations conducted
over the last four decades (e.g. Goldreich & Tremaine 1979, 1980;
Lin & Papaloizou 1986a,b; Paardekooper & Mellema 2004, 2006).
The fundamental question is whether or not these structures are cre-
ated by embedded protoplanets (e.g. Flock et al. 2015; Zhang et al.
2015; Dipierro et al. 2015; Dong et al. 2015; Jin et al. 2016).
However, most of the theory on gaps in discs has consid-
ered purely gaseous discs, with dust responding to the gas distri-
bution mainly via drag. As a result, there is one physical mech-
anism discussed in the literature for explaining the origin of
planet-induced gaps and ring-like structures in dusty discs: High
mass planets exert a tidal torque that overpowers the local vis-
cous torque and forms a gap in the gas (Papaloizou & Lin 1984;
Bryden et al. 1999). Large grains then drift towards the pressure
maxima located at the gap edges as a result of gas drag. Trap-
ping of dust at the pressure maxima prevents the large grains
from being accreted onto the planet and produces a deep structure
in the dust (Paardekooper & Mellema 2004, 2006; Fouchet et al.
2007, 2010), easily detectable by ALMA (Pinilla et al. 2012,
2015; Gonzalez et al. 2012, 2015). Small grains, however, remain
strongly coupled to the viscous flow of the gas and are accreted
onto the planet (Rice. et al. 2006).
Here, motivated by our recent numerical two fluid dust/gas
simulations modelling the HL Tau disc (Dipierro et al. 2015),
⋆ [email protected]
© 2016 The Authors
we point out that the standard model only partially captures the
physics of gap opening in dusty discs. We use our 3D dust-and-gas
Smoothed Particle Hydrodynamics (SPH) code to demonstrate the
different roles of the tidal and drag torques in opening dust gaps in
protoplanetary discs. In particular, we identify regimes for which it
is possible for gaps to form only in the dust, but not in the gas.
2 METHODOLOGY
2.1 Dust and gas simulations
We use the PHANTOM SPH code (Price & Federrath 2010;
Lodato & Price 2010; Price 2012; Nixon, King & Price 2013) to
perform global 3D simulations of dusty discs containing one or
more embedded planets. The three processes involved in the gap-
opening process -- gas viscosity, gravity and drag -- are computed
consistently. Artificial viscosity is applied, calibrated to mimic the
viscous transport of gas described with a Prandtl-like model of tur-
bulence (Lodato & Price 2010). The central star and the planets are
modelled using sink particles. Sink particles are allowed to both
accrete and migrate as a consequence of their interactions with the
disc (Bate, Bonnell & Price 1995) and their gravitational interac-
tion. In particular the gravity between the sink particles and the gas
and dust particles in the disc is modelled.
Drag between gas and dust is computed using the two-fluid al-
gorithm described in Laibe & Price (2012). The algorithm has been
extensively benchmarked on test problems with known analytic so-
lutions (Laibe & Price 2011, 2012; Price & Laibe 2015). We addi-
tionally perform simulations where gravity from the planets is arti-
ficially turned off in the dust. This allows us to contrast our results
with a model where the dust is assumed to respond to the gas only
via the drag torque.
2
Dipierro et al.
gas
0.05 MJ
0.1 MJ
-2
-1.5
-1
-4 -3.5 -3 -2.5 -2
Figure 1. Gap opening via Mechanism I, where low mass planets carve a gap in the dust but not the gas. Plots show gas (left) and millimetre dust grain (centre)
surface densities in a dusty disc hosting planets of mass 0.05 MJ (top row) and 0.1 MJ (bottom row). While the 0.05 MJ creates a depletion of dust at the
planet location, the 0.1 MJ planet is able to carve a gap in the dust. Neglecting the gravity between the planet and the dust (right panels) shows that the gap is
opened by the tidal torque. The drag torque acts to close the gap due to the radial migration of dust particles from the outer disc (top centre panel).
2.2 Initial conditions
3 RESULTS
We setup a disc as in Lodato & Price (2010). We assume a central
star of mass 1.3 M⊙ surrounded by a gas disc made of 5 × 105 gas
particles and a dust disc made of 3 × 105 dust particles. The two
discs extend from rin = 1 au to rout = 120 au. We model the initial
surface density profiles of the discs using power-laws of the form
Σ(r) = Σin(r/rin)−p. We adopt p = 0.1 and set Σin such that the
total gas mass contained between rin and rout is 0.0002 M⊙. We
assume 1 mm dust grains with a corresponding Stokes number (the
ratio between the stopping time and the orbital timescale), St ∼ 10.
The initial dust-to-gas ratio is 0.01 and St ∝ 1/Σg ∼ r0.1 in the
disc. We simulate only the inner part of the disc since this is what
can be observed with ALMA e.g. in HL Tau. If the gas phase were
to extend to rout = 1000 au, the total mass of the system is ≃ 0.01
M⊙. We assume a vertically isothermal equation of state P = c2
s ρ
with cs(r) = cs,in(r/rin)−0.35 and an aspect ratio of the disc that
is 0.05 at 1 au. We set an SPH viscosity parameter αAV = 0.1 giv-
ing an effective Shakura & Sunyaev (1973) viscosity αSS ≈ 0.004.
We setup a planet located at 40 au and evolve the simulations over
40 planetary orbits. This is sufficient to study the physics of dust
gap opening with our assumed grain size, though we caution that
further evolution occurs over longer timescales. We vary the planet
mass in the range [0.05, 0.1, 0.5, 1] MJ to evaluate the relative con-
tributions of the tidal and drag torques.
Gap formation is a competition between torques. In a gas disc the
competition is between the tidal torque from the planet trying to
open a gap and the viscous torque trying to close it. Dust, by con-
trast, is pressureless and inviscid, and the competition is between
the tidal torque and the aerodynamic drag torque.
Dust efficiently settles to the midplane in our simulations,
forming a stable dust layer with dust to gas scale height ratio of
∼ pαSS/St ∼ 0.02, consistent with the Dubrulle et al. (1995)
model and other SPH simulations of dusty discs (e.g. Laibe et al.
2008). Settling of grains is expected to slightly reinforce the con-
tribution from the tidal torque by local geometric effects.
3.1 Mechanism I -- low mass planets
Fig. 1 demonstrates gap-opening when the planet is not massive
enough to carve a gap in the gas disc. The gas shows only a weak
one-armed spiral density wake supported by pressure, as predicted
by linear density wave theory (Ogilvie & Lubow 2002).
The general expression for the drag torque is
Λd = −r
K
ρd
(vφ
d
− vφ
g ),
(1)
where K is the drag coefficient, ρd is the dust density and vφ
d
MNRAS 000, 1 -- 5 (2016)
and vφ
g are the azimuthal velocity of the dust and gas, respec-
tively. For large grains and in stationary regime, the dust and the
gas velocities are given by eq. 2.12-2.14 of Nakagawa et al. (1986)
(assuming corrections due to gas viscosity are negligible). Hence
Λd ∝ (St+St−1)−1 since it is dominated by the contribution from
the background radial pressure gradient (Nakagawa et al. 1986). In
a disc undisturbed by a planet the radial pressure gradient and the
gas viscous velocity are, in general, negative. Hence the drag torque
on the dust phase is negative, pushing dust particles inwards. The
planet exerts a tidal torque on the dust which is positive outside
the planet orbit and negative inside, forcing grains away from the
planet. This torque increases as the planet mass increases. The bal-
ance between these two torques determines the evolution of the
dust: A gap opening process resisted by drag from the outer disc.
At radii smaller than the orbital radius of the planet, both con-
tributions add to make the grains drift inwards. At larger radii the
tidal torque resists the drag torque by repelling the particles from
the planet. For any planet mass, the tidal torque therefore decel-
erates dust particles outside the planetary orbit. A strong enough
tidal torque in the outer disc prevents the replenishment of the dust
population in the inner disc due to the action of the drag torque.
Gap closing (opening) induced by drag outside (inside) the plan-
etary orbit is maximal for grains with St = 1, decreases linearly
with Stokes number for St ≪ 1 and is inversely proportional to
Stokes number for St ≫ 1 (Takeuchi & Artymowicz 2001). How-
ever, small grains follow the viscous evolution of the gas. We there-
fore expect a less effective replenishment of larger particles from
outside the planetary orbit due to the reduced drag torque, resulting
in a more symmetric gap around the planet location.
Whether or not a gap can be opened in the dust depends on
the planet mass. For a planet of very low mass (top row in Fig. 1;
0.05 MJ), we only see a partial depletion of dust at the planet lo-
cation, while the regions of the disc inside the planetary orbit are
constantly replenished by an inflow of particles from the outer disc.
Replenishment occurs because the drag torque produces a constant
inward drift of dust particles, while the action of the the tidal torque
is localised at the planet position. If the time taken by dust particles
to refill the gap is shorter than the orbital timescale, the gap (and
inner disc) is refilled by the action of the drag torque. However, if
the tidal torque from the planet on the inner disc is strong enough
to overcome this refilling, an inner cavity will slowly open.
The lower panel in Fig. 1 shows that a sufficiently massive
planet (here 0.1 MJ) clears a complete orbit in the dust. This oc-
curs once the tidal torque is large enough to overcome the drag
torque outside of the planetary orbit. Fig. 2 shows that the gap is
not centred around the planet orbit, but is shifted towards the inner
disc. Initially the gap is symmetric around the planet location, but
becomes more and more asymmetric as the dust in the inner part of
the disc drifts inwards, away from the planets orbit. As a result, no
dust population is maintained in the corotation region. We confirm
our physical picture above by noting that our 0.05 MJ simulation
shows the same gap structure after a further 20 orbits, while the gap
in the 0.1 MJ simulation grows more asymmetric with time.
The right panel of Fig. 1 demonstrates the role played by the
tidal torque in the gap-opening mechanism, showing the dust dy-
namics computed with the gravitational force from the planet on the
dust phase switched off. In this case, solid particles are unaffected
by the surrounding gas and simply migrate towards the central star.
MNRAS 000, 1 -- 5 (2016)
Gap opening in dusty discs
3
Figure 2. Evolution of the azimuthally averaged dust surface density profile
for the disc hosting a 0.1 MJ planet corresponding to the bottom centre
panel of Fig. 1. The dotted-dashed line provides the gas surface density
scaled by a factor of 0.01 at the end of the simulation, for direct comparison
with the dust phase. The dotted vertical line indicates the planet orbit. An
asymmetric gap profile is seen in the dust, but not in the gas.
3.2 Mechanism II -- high mass planets
Previous authors have identified two regimes for dust gap for-
mation: high-mass planets opening gas and dust gaps (e.g.
Crida et al. 2006) and low-mass planets creating mild radial pres-
sure gradients in the gas but deeper gaps in the dust (e.g.
Paardekooper & Mellema 2004). Both of these regimes involve the
same, drag assisted, gap opening mechanism (our Mechanism II),
but with different intensities. For sufficiently massive planets (0.5
MJ and 1 MJ, top and bottom rows, respectively), the tidal torque
dominates the viscous torque at the planet location, expelling the
gas. This results in a stable gap in the gas surface density. A pres-
sure maximum is always created at the outer edge of the gap.
At the inner edge, a pressure maximum forms only if the per-
turbation of the pressure profile induced by the tidal torque ex-
ceeds the background pressure gradient. This is more likely to oc-
cur for flat pressure profiles and a planet located close to the star
(Crida & Morbidelli 2007; Fouchet et al. 2010). The drag torque on
the dust phase is influenced by the formation of these pressure max-
ima at the gap edges. Dust accumulates at the location of the pres-
sure maxima, forming a deeper dust gap than in the gas.
As expected, the width and depth of the gaps increase as the
planet mass increases (top and bottom rows in Fig. 3). Compared
to the low mass planet case, the additional pile-up induced by the
presence of pressure maxima concentrates dust more efficiently at
the outer edges of the gap. Fig. 4 shows that the gaps in the dust disc
are W-shaped and asymmetric around the location of the planet.
Dust grains located initially between the inner edge of the gap in
the gas and the corotation radius are depleted towards the inner
pressure maximum by the drag torque. In contrast to Mechanism I,
particles with St = 1 concentrate fastest on the gap edges. It is
therefore easier for marginally coupled particles to open gaps in
the dust (Weidenschilling 1977).
A large and stable population of dust grains is observed in the
corotation region where the tidal torque provides a stabilising ef-
fect. The efficiency of the drag torque decreases locally as the gas
surface density decreases (Laibe et al. 2012) and the background
pressure gradient is significantly reduced (see left panel of Fig. 4).
Interestingly, excitation of particles eccentricities at the outer edge
4
Dipierro et al.
gas
dust
0.5 MJ
1 MJ
-2
-1.5
-1
-5
-4
-3
-2
Figure 3. Gap opening via Mechanism II, where high mass planets carve a partial or total gap in the gas, and the dust is evacuated from the gap via drag and
tidal torques. Plots are as in Fig. 1, but with planet masses 0.5 MJ (top) and 1 MJ (bottom). Although the tidal torque modifies the structure of the gap and
stabilises the corotation region, the structure in the dust phase is dominated by the drag torque (comparing centre and right panels).
10−1
]
2
−
m
c
g
[
g
Σ
Mp = 0.5 MJ
Mp = 1 MJ
10−2
10−3
10−4
10−5
]
2
−
m
c
g
[
d
Σ
Mp = 0.5 MJ
Mp = 1 MJ
Mp = 1 MJ
grav + drag
drag
10−2
0
20
40
60
r [au]
80
100
120
10−6
0
20
40
60
r [au]
80
100
120
0
20
40
60
r [au]
80
100
120
Figure 4. Azimuthally averaged surface density of the gas corresponding to the simulations of Fig 3. The dotted vertical line indicates the planet's location.
Gaps are created in both the gas and the dust phases.
of the gap forms narrow ridges just outside the orbit of the planet
(Ayliffe et al. 2012; Picogna & Kley 2015). This effect does not oc-
cur inside of the orbital radius of the planet because the drag torque
is strong enough to efficiently damp any resonances that develop
(Fouchet et al. 2007; Ayliffe et al. 2012), similar to what occurs in
the whole disc in Mechanism I. For our 1 MJ planet, the outer edge
of the dust gap is close to the 3:2 resonance (r ∼ 52 au), inducing
a double peaked outer edge in the dust density profile (see Fig. 4).
The right panel of Figs 3 and 4 shows that for high mass plan-
ets, the formation of a gap in the dust can be recovered simply by
considering drag effects and neglecting the action of the gravita-
tional potential of the planet. However, the detailed structure of the
gap is still different when the tidal torque is included: the gap is
wider, deeper, with a corotation region, sharper edges and more
asymmetries due to external resonances.
MNRAS 000, 1 -- 5 (2016)
gas
dust
Gap opening in dusty discs
5
-2.5 -2 -1.5 -1 -0.5
-3.5 -3 -2.5 -2 -1.5
Figure 5. Similar to Fig. 1, but with a disc hosting three planets of mass 0.08, 0.1 and 0.52 MJ initially located at the same distance as the gaps detected in
HL Tau. The first two planets open dust gaps according to Mechanism I, whereas the third planet is at the transition between the two mechanisms. Importantly,
the tidal torque in the dust can not be neglected.
3.3 Application to the HL Tau disc
REFERENCES
Based on the previous discussion, we can now interpret our mod-
elling of HL Tau (Dipierro et al. 2015) in the context of the two
different mechanisms for dust gap opening. In this model, we used
a dust-to-gas ratio of ∼ 0.06 to reproduce a n(s) ∝ s−3.5 grain
size distribution. We setup three planets located at 13.2, 32.3 and
68.8 au, with masses 0.08, 0.1 and 0.52 MJ, and accretion radii
0.25, 0.25 and 0.75 au respectively. Fig. 5 shows the gas and the
dust surface densities after 10 orbits of the third planet. The first
two planets open gaps via Mechanism I, with gaps in dust but not
gas. The third planet has a mass at the transition between the two
mechanisms, chaning the gas density slightly and carving a local
gap which is rapidly filled by the viscous inflow of gas. Hence, we
can interpret the gaps observed by ALMA with these two mecha-
nisms (although alternative scenarios may work as well).
4 CONCLUSION
We have identified two physical mechanisms for dust gap opening
by embedded planets in dusty protoplanetary discs. Our conclu-
sions are:
i) Gap formation in the gas is not a necessary condition for open-
ing a gap in the dust.
ii) For low mass planets that do not produce pressure maxima
in the gas, drag resists (assists) the tidal torque outside (inside) the
planetary orbit, forming an asymmetric gap around the planet orbit.
iii) For high mass planets that create pressure maxima in the gas,
solid particles are prevented from accreting onto the planet. Here
drag assists the tidal torque, leading to the formation of a deep dust
gap with a stable population of grains at the corotation region.
ACKNOWLEDGMENTS
We thank the referee for insightful comments. GD thanks Monash
for CPU time on NeCTaR. We acknowledge an ARC Future Fel-
lowship and Discovery Project. GD and G. Lodato acknowledge
funding via PRIN MIUR prot. 2010LY5N2T. G. Laibe is funded
by ERC FP7 grant ECOGAL. We used SPLASH (Price 2007).
MNRAS 000, 1 -- 5 (2016)
ALMA Partnership et al., 2015, ApJ, 808, L3
Ayliffe B. A., Laibe G., Price D. J., Bate M. R., 2012, MNRAS, 423, 1450
Bate M. R., Bonnell I. A., Price N. M., 1995, MNRAS, 277, 362
Bryden G., et al., 1999, ApJ, 514, 344
Crida A., Morbidelli A., 2007, MNRAS, 377, 1324
Crida A., Morbidelli A., Masset F., 2006, Icarus, 181, 587
Dipierro G., et al., 2015, MNRAS, 453, L73
Dong R., Zhu Z., Whitney B., 2015, ApJ, 809, 93
Dubrulle B., Morfill G., Sterzik M., 1995, Icarus, 114, 237
Flock M., et al., 2015, A&A, 574, A68
Fouchet L., et al., 2007, A&A, 474, 1037
Fouchet L., Gonzalez J.-F., Maddison S. T., 2010, A&A, 518, A16
Goldreich P., Tremaine S., 1979, ApJ, 233, 857
Goldreich P., Tremaine S., 1980, ApJ, 241, 425
Gonzalez J.-F., Pinte C., Maddison S. T., M´enard F., 2012, A&A, 547
Gonzalez J.-F., et al., 2015, MNRAS, 454, L36
Jin S., Li S., Isella A., Li H., Ji J., 2016, preprint, (arXiv:1601.00358)
Laibe G., Price D. J., 2011, MNRAS, 418, 1491
Laibe G., Price D., 2012, MNRAS, 420, 2345
Laibe G., Gonzalez J.-F., Fouchet L., Maddison S. T., 2008, A&A, 487, 265
Laibe G., Gonzalez J.-F., Maddison S. T., 2012, A&A, 537, A61
Lin D. N. C., Papaloizou J., 1986a, ApJ, 307, 395
Lin D. N. C., Papaloizou J., 1986b, ApJ, 309, 846
Lodato G., Price D. J., 2010, MNRAS, 405, 1212
Nakagawa Y., Sekiya M., Hayashi C., 1986, Icarus, 67, 375
Nixon C., King A., Price D., 2013, MNRAS, 434, 1946
Nomura H., et al., 2015, preprint, (arXiv:1512.05440)
Ogilvie G. I., Lubow S. H., 2002, MNRAS, 330, 950
Paardekooper S.-J., Mellema G., 2004, A&A, 425, L9
Paardekooper S.-J., Mellema G., 2006, A&A, 453, 1129
Papaloizou J., Lin D. N. C., 1984, ApJ, 285, 818
Picogna G., Kley W., 2015, A&A, 584, A110
Pinilla P., Benisty M., Birnstiel T., 2012, A&A, 545
Pinilla P., et al., 2015, A&A, 573
Price D. J., 2007, Publ. Astron. Soc. Australia, 24, 159
Price D. J., 2012, Journal of Computational Physics, 231, 759
Price D. J., Federrath C., 2010, MNRAS, 406, 1659
Price D. J., Laibe G., 2015, MNRAS, 451, 813
Rice. W. K. M., et al., 2006, MNRAS, 373, 1619
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Takeuchi T., Artymowicz P., 2001, ApJ, 557, 990
Weidenschilling S. J., 1977, MNRAS, 180, 57
Zhang K., et al., 2015, ApJ, 806, L7
|
1806.09746 | 1 | 1806 | 2018-06-26T01:07:54 | Dust Production Rates in the Fomalhaut Debris Disk from SOFIA/FORCAST Mid-infrared Imaging | [
"astro-ph.EP"
] | We present the first spatially resolved mid-infrared (37.1 $\mu$m) image of the Fomalhaut debris disk. We use PSF fitting and subtraction to distinctly measure the flux from the unresolved component and the debris disk. We measure an infrared excess in the point source of $0.9 \pm 0.2$ Jy, consistent with emission from warm dust in an inner disk structure (Su et al. 2016), and inconsistent with a stellar wind origin. We cannot confirm or rule out the presence of a pileup ring (Su et al. 2016) near the star. In the cold region, the 37 $\mu$m imaging is sensitive to emission from small, blowout grains, which is an excellent probe of the dust production rate from planetesimal collisions. Under the assumptions that the dust grains are icy aggregates and the debris disk is in steady state, this result is consistent with the dust production rates predicted by Kenyon & Bromley (2008) from theoretical models of icy planet formation. We find a dust luminosity of $(7.9 \pm 0.8) \times 10^{-4}$ L$_\odot$ and a dust mass of 8 -- 16 lunar masses, depending on grain porosity, with $\sim 1$ lunar mass in grains with radius 1 $\mu$m -- 1 mm. If the grains are icy and highly porous, meter-sized objects must be invoked to explain the far-IR, submm, and mm emission. If the grains are composed of astronomical silicates, there is a dearth of blowout grains (Pawellek et al. 2014) and the mass loss rate is well below the predicted dust production values. | astro-ph.EP | astro-ph | Submitted to ApJ
Preprint typeset using LATEX style AASTeX6 v. 1.0
8
1
0
2
n
u
J
6
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
4
7
9
0
.
6
0
8
1
:
v
i
X
r
a
DUST PRODUCTION RATES IN THE FOMALHAUT DEBRIS DISK FROM SOFIA/FORCAST
MID-INFRARED IMAGING
J. D. Adams1, T. L. Herter2, R. M. Lau3, C. Trinh1, M. Hankins2
(Received; Revised; Accepted)
1Stratospheric Observatory for Infrared Astronomy, Universities Space Research Association, NASA/Armstrong Flight Research Center,
2825 East Avenue P, Palmdale, CA 93550 USA
2Department of Astronomy, Cornell University, Space Sciences Bldg., Ithaca, NY 14853 USA
3Division of Physics, Mathematics and Astronomy, California Institute of Technology, 1200 E. California Blvd., Pasadena, CA 91125 USA
ABSTRACT
We present the first spatially resolved mid-infrared (37.1 µm) image of the Fomalhaut debris disk. We
use PSF fitting and subtraction to distinctly measure the flux from the unresolved component and
the debris disk. We measure an infrared excess in the point source of 0.9 ± 0.2 Jy, consistent with
emission from warm dust in an inner disk structure (Su et al. 2016), and inconsistent with a stellar
wind origin. We cannot confirm or rule out the presence of a pileup ring (Su et al. 2016) near the
star. In the cold region, the 37 µm imaging is sensitive to emission from small, blowout grains, which
is an excellent probe of the dust production rate from planetesimal collisions. Under the assumptions
that the dust grains are icy aggregates and the debris disk is in steady state, this result is consistent
with the dust production rates predicted by Kenyon & Bromley (2008) from theoretical models of icy
planet formation. We find a dust luminosity of (7.9 ± 0.8) × 10−4 L⊙ and a dust mass of 8 – 16 lunar
masses, depending on grain porosity, with ∼ 1 lunar mass in grains with radius 1 µm – 1 mm. If the
grains are icy and highly porous, meter-sized objects must be invoked to explain the far-IR, submm,
and mm emission. If the grains are composed of astronomical silicates, there is a dearth of blowout
grains (Pawellek et al. 2014) and the mass loss rate is well below the predicted dust production values.
Keywords: infrared: planetary systems - stars: circumstellar matter – planetdisk interactions
1. INTRODUCTION
Debris disks around stars are thought to be formed from a collisional cascade that occurs when planets form and
stir planetesimals. Such collisions continually produce dust around the star. Smaller dust grains are blown out of
the system by radiation pressure, while larger grains are pulled into the inner regions of the system by the Poynting-
Robertson effect. However, the dust is replenished by the ongoing collisions which occur over the lifetime of the star.
Kenyon & Bromley (2008) have shown that the range of dust production varies with stellar mass, primordial disk mass,
and stellar age. As protoplanets form around 1 – 3 M⊙ stars, dust production rises rapidly over 1 – 10 Myr, reaches a
peak, then declines over the lifetime of the star. The peak dust production rates for 1 – 3 M⊙ ranges from ∼ 1019 to
∼ 1022 g yr−1. More massive disks exhibit peak dust production earlier (∼ 3 Myr) than less massive disks (∼ 10 Myr).
For 1 – 3 M⊙ stars with ages > 30 Myr, the range of predicted dust production rates narrows and becomes verifiable
with observations. Specifically, the detection of small grains, coupled with calculations of their blowout timescales,
can be used to determine the dust production rate in a debris disk that is in steady state.
Fomalhaut (α PsA) is a nearby (7.7 pc, van Leeuwen 2007), young (440 Myr, Mamajek 2012) A3V star with an
infrared excess and an active debris disk containing a cold submillimeter ring (Holland et al. 2003) that is interpreted
as a Kuiper Belt analog. It also has a candidate planetary object, Fomalhaut b (Kalas et al. 2013). The disk was
first discovered as infrared excess in IRAS observations (Aumann 1985). The infrared excess is known to originate
very close to the star, as seen in Spitzer/IRS spectra (Stapelfeldt et al. 2004) and near-IR interferometric imaging
(Absil et al. 2009). This excess has been interpreted as having a stellar wind origin (Acke et al. 2012) or as signature
of hot dust in the inner disk. Lebreton et al. (2013) proposed that the infrared excess arises from a dust-producing belt
at 1 – 2 AU from the star, with an inner hot ring of dust located at 0.1 – 0.3 AU. However, this model overpredicts the
mid-IR flux (Su et al. 2016). Su et al. (2013) suggested that the dust arises from an asteroid belt, with the near-IR
excess originating in hot, magnetically trapped nano grains near the star. The proposed asteroid belt lies near the
2
Adams et al.
ice-line at 8 – 15 AU (Su et al. 2016), resulting in the production of dust, a drag-in disk from the Poynting-Robertson
effect, and a pileup ring due to density enhancement near the silicate sublimation radius.
The cold belt is well-defined (Acke et al. 2012, MacGregor et al. 2017) and exhibits apocenter glow (MacGregor et al.
2017) due to the eccentricity of the system. The disk has been observed in the mid-infrared by Stapelfeldt et al. (2004),
who used Spitzer/MIPS to obtain images of the disk at 24, 70, and 160 µm. At 24 µm, the ansai of the disk were
detected; however, the Spitzer observations lacked the spatial resolution to completely resolve the disk from the star.
Recently, Acke et al. (2012) used Herschel/PACS (Poglitsch et al. 2010) and Herschel/SPIRE (Griffin et al. 2010) to
image the disk in broad passbands at 70 – 500 µm. Although the disk is resolved at 70 µm, the Herschel data are
sensitive to large grains, and are less sensitive to the small grain population.
An estimate of the dust production rate requires knowledge of the dust grain composition. Fomalhaut shows an
asymmetry in scattered light (Kalas, Graham, & Clampin 2005), which could be interpreted as backscattering from
large grains (Min et al. 2010). Fluffy aggregates have been considered to explain the observed scattering and thermal
emission components to the dust SED (Acke et al. 2012). These grains are similar to icy, porous cometary debris
seen in the solar system (Fraundorf et al. 1982). Using fluffy aggregates, Acke et al. (2012) derived a mass loss rate
of 2 × 1021 g yr−1 for 1 – 13 µm grains in the cold component. However, this value is uncertain due to sparse data in
the mid-IR.
An alternate interpretation of the Herschel data was present by Pawellek et al. (2014), who used Herschel data and
astronomical silicate (Draine 2003) models to examine the grain size distribution in the spatially resolved component.
No grains smaller than the blowout size were found, indicating that, if the grains are silicates, the smallest grain size
formed in collisions is larger than the blowout size or the grains are rapidly blown out. These grains are iceless despite
their location outside the ice sublimation zone.
Motivated to detect small grains and determine a precise measurement of the dust production rate, we obtained new
mid-infrared observations with SOFIA (Young et al. 2012). In §2 and §3, we discuss SOFIA observations and results.
In §4, we discuss dust models and results based on icy aggregate and astronomical silicate compositions. In §5, we
derive the mass loss rates as a result of stellar radiation pressure. §6 contains a discussion of the results for both the
point source and the Kuiper Belt analog. Our conclusions are presented in §7.
2. OBSERVATIONS
We observed Fomalhaut with the FORCAST instrument (Herter et al. 2012) on SOFIA at a wavelength of 37.1
µm (direct imaging mode), under SOFIA Program 04 0064. Observations were carried out over ∼ 10 hours on four
flights during Observing Campaign OC4-G (Southern Hemisphere deployment) in July 2016 (Table 1). FORCAST
has a field-of-view of 3.4′ × 3.2′. The observations were performed with matched chopping and nodding (NMC mode)
and dithering over a grid of nine positions spaced at 30′′. The off-source fields were chosen to be orthogonal to the
apparent major axis of the disk in order to minimize the chop throw, thereby minimizing the amount of coma in the
telescope beam. Typical dwell times in each nod beam were approximately 60 s, including chopping inefficiencies. A
small number of observations were identified to have been taken under aircraft door vignetting, which occurs when
the telescope is near its elevation limit and causes significant background variations.
The raw data were processed using the pipeline described in Herter et al. (2013). This pipeline performs chop and
nod subtraction and applies corrections for droop, detector non-linearity, multiplexer crosstalk, and optical distortion.
The chop, nod, and dither positions were aligned using the centroid of the star as a reference position. We averaged
the aligned images (excluding the vignetted data), producing an image with an effective on-source exposure times of
24,302 s. The combined signals were individually weighted to account for the difference in signal between the matched
beams and the non-matched beams. Finally, we subtracted residual background structure across the image caused by
the presence of a bright central source (Lau et al. 2013) by using clipping and a second order polynomial fit to the
background across each row. The rectified plate scale for the final image was 0.768′′ per pixel. The effective RMS
noise level was ∼ 0.8 mJy/pixel.
Calibration factors at each level altitude were derived from the average photometric instrument response to standard
stars observed during the observing campaign, with corrections for elevation angle and altitude. The calibration factors
are for a flat spectrum source. The calibration uncertainty was ∼ 6% (Table 1). We estimate a total uncertainty of
∼ 10% resulting from both calibration and flat fielding errors.
3. OBSERVATIONAL RESULTS
3.1. Imagery
Dust Production Rates in the Fomalhaut Debris Disk
3
In Figure 1 (left panel), we show the processed 37 µm image of Fomalhaut. The signal-to-noise ratio per pixel is
low (2) and therefore we have to rely on integrated signal for flux measurements. In Figure 1 (right panel), we show
the processed 37.1 µm image convolved with a Gaussian kernel to enhance the signal-to-noise ratio. The Gaussian
kernel had a FWHM of 6.2′′ × 6.2′′. The disk ansai are visible, with average signal-to-noise ratios of 48 per pixel in the
Southeastern ansa and an average signal-to-noise ratio of 25 in the Northwestern ansa. For the subsequent analysis,
we use unconvolved data for PSF subtraction and to derive the disk flux.
3.2. PSF subtraction
In order to subtract the unresolved contribution from the disk flux, we performed PSF fitting on a standard star and
the Fomalhaut point source. We constructed a model for the PSF consisting of a modified Airy function for diffraction,
a polynomial component in the radial direction, and an elliptical Gaussian function for telescope jitter. The source of
the a polynomial component is unclear, but it is prominent in the first Airy bright ring. A similar approach was used
by Su et al. (2017), who fitted the FORCAST PSF at 35 µm using a hybrid PSF based on a model and empirical,
standard object PSFs. The model parameters were first adjusted to fit the radial profile of a standard star. We
used observations of α Boo at 37.1 µm, taken during OC4-G, as a standard star reference for PSF fitting. The RMS
jitter values were 0.′′.84 and 0.′′55 in the cross-elevation and elevation directions, respectively. An Airy function for
diffraction was modified by a third order polynomial across the first bright ring. To enable us to subtract the PSF
from the Fomalhaut image, we slightly adjusted to fit the radial profile of the Fomalhaut point source to account for
variability in the PSF with telescope elevation. For Fomalhaut, the RMS jitter values were 1.′′.1 and 0.′′84 in the
cross-elevation and elevation directions, respectively
Figure 2 shows an image of α Boo, a radial profile of α Boo and the model fit to the data. Figure 2 also shows
the image and residuals after the model has been subtracted from the data. The scatter in residuals is comparable
to the background noise, indicating clean subtraction. Figure 3 shows analogous PSF fitting and subtraction for our
Fomalhaut data. The PSF-subtracted image clearly shows the belt in the cold region.
3.3. Flux profile
We generated a flux profile along the apparent major axis of the disk by summing the flux in the direction per-
pendicular to the apparent major axis. The width of the summed region was 23.′′9 and the position angle of the
apparent major axis was 156◦ (E of N). Figure 4 shows this flux profile along the apparent major axis both with and
without the PSF subtracted. The cold belt is spatially unresolved in the radial direction due to its narrow width
(13AU, MacGregor et al. 2017). The outer edge of the Southwestern ansa is located approximately 17′′ (131 AU) and
the outer edge of the Northwestern ansa is located approximately 19.′′3 (148 AU) from the star (Figure 4). Taking
into account blurring by the PSF, these results agree well with the semi-major axis of the outer edge of the belt as
measured by ALMA (149 ± 2 AU, MacGregor et al. 2017). The Southeastern section shows emission from an inner
disk. A fainter flux level in the northwestern section precluded detection of the inner disk. A faint halo is seen in
the flux profile extending ∼ 10′′ beyond the edges of the ansai, but it is only marginally detected by FORCAST. In
archival Herschel images1 at 70 µm, this halo extends to a radius of ∼ 60′′ (460 AU).
3.4. Integrated fluxes
We used a tilted annulus to sum the flux over the region containing the Kuiper Belt analog and its halo. The annuli
covered the 100 – 450 AU (13′′ – 58′′) region to account for blurring by the PSF and the extent of the halo as seen
in archival 70 µm Herschel images. The inclination (65.6◦), position angle (156◦), and eccentricity (0.12) values were
taken from Acke et al. (2012) and MacGregor et al. (2017). The average background per pixel was measured in a
tilted annulus with radius 58′′ − 60′′ and subtracted from each coadded pixel. We measured an integrated flux density
of 2.0 ± 0.2 Jy in the cold region.
To compute the flux of the point source, we integrated the fitted model PSF. We applied a 15% color correction for
the contribution from the photosphere. The photospheric flux density is expected to be ∼ 1.15 Jy (Castelli & Kurucz
2004). The color correction for unresolved warm dust is negligible. Our final result is 2.25 ± 0.2 Jy for a flat spectrum
source, which is color-corrected to 2.1 ± 0.2 Jy for the point source. This result yields an infrared excess of 0.9 ± 0.2
Jy at 37.1 µm.
4. DUST MODELING
1 http://http://archives.esac.esa.int/hsa/whsa/
4
Adams et al.
In order to characterize the dust populations in the debris disk, we performed dust modeling using a radiation field
generated by the star and the assumption that the dust in the disk is in thermodynamic equilibrium. The luminosity
Lg of a single grain of radius a is given by:
Lg = πa2 R2
∗
r2 Z F (λ)Q(λ, a)dλ = 4π2a2Z B(λ, T )Q(λ, a)dλ
(1)
where R∗ is the stellar radius, r is the distance from the star to the dust grain, F (λ) is the stellar flux density at
the surface of the star, Q(λ, a) is the ratio of the absorption area to the geometric cross-sectional area, and T is the
equilibrium temperature of the dust grain.
4.1. Radiation field
For the stellar flux density, we used a model stellar atmosphere with an effective temperature of 8,750 K and solar
metallicity from Castelli & Kurucz (2004). We used an average distance from the star of 140 AU for the cold region.
The intervening material is optically thin based on its transparency in visible light images (Kalas, Graham, & Clampin
2005).
4.2. Grain composition and size distributions
For the grain composition, we adopted icy aggregate grains as suggested for Fomalhaut by Acke et al. (2012).
We considered a range of porosity, from compact grains to highly porous grains such as those seen in AU Mic
(Graham, Kalas & Matthews 2007). The icy grains are composed by mass of 22.5% silicates, 30.1% organics, and
47.4% water ice (Pollack et al. 1994, Kataoka et al. 2014), with a mean bulk density of 1.68 g cm−3. Q(λ, a) was
computed analytically from the real and imaginary refractive indices and derivations given in Kataoka et al. (2014)
for volume filling factors f of 1.0, 0.5, 0.1, and 0.05.
We also considered astronomical silicate grains as the source of dust emission (Draine 2003). They are compact,
with a bulk density of 3.3 g cm−3 and a blowout radius of ∼ 5 µm for Fomalhaut.
The grain size distribution N (a) was modeled as dN/da ∝ aq. The size distributions for blowout grains and for
grains larger than the blowout size evolve differently, so we used a separate size distributions for the cases a ≤ ablow
and a ≥ ablow. The limits on the grain sizes are set by amin and amax, the minimum and maximum grain sizes,
respectively. For very large icy grains (2πa/λ ≫ 1), the absorption coefficients approach 1 − 0.1f and the scattering
coefficients approach 1 + 0.1f in the mid- and far-IR. q, amin, amax are the only free parameters in the dust model.
The number of dust grains N (a) and integrated luminosity of the dust populations are constrained by the observed
SED.
4.3. Modeling Results
For icy grains in the cold region, we computed equilibrium temperatures that fall in the range 48 – 69 K. Fig.
5 shows equilibrium temperature T (a) as a function of grain radius a. The grains in the halo will be at a slightly
lower temperature due to their larger distance from the star than the cold belt at a distance of ∼ 140 AU, but their
contribution to the integrated flux is small.
Figure 6 shows the SED for the cold region, including data from Spitzer/MIPS at 24 µm (Stapelfeldt et al. 2004)
and outer disk fluxes at 70 – 500 µm from Acke et al. (2012). We note that the outer disk fluxes at 70 – 500 µm from
Acke et al. (2012) were derived from a dust model that was convolved with the Herschel beams and fit to the Herschel
images. Such an approach relies on grain composition and density distribution to separate the flux of the outer disk
from that of the inner disk. Figure 6 also shows submillimeter data at 350 µm (Marsh et al. 2005) and 450 and 850
µm (Holland et al. 2003), as well as the integrated flux at 1.3 mm from ALMA (MacGregor et al. 2017).
Icy grain model SEDs were fitted to the observed SED by eye (χ2 ≈ 0.02) for volume filling factors f of 1.0, 0.5,
0.1, and 0.05. Table 2 summarizes the best-fit model parameters and corresponding results for the cold belt. High
porosity grains require very large maximum grain sizes, up to 1.2 m for f = 0.05. Figure 6 shows the SEDs for the
model cases corresponding to the four values of volume filling factors. All these cases can reproduce the observed
SED. Consequently, the existing data set does not effectively constrain porosity. Integrating the model SEDs yields
the dust luminosities, while integrating the size distribution and grain masses yields the dust mass. The dust mass
values are listed in Table 2. The dust luminosity is tightly constrained at (7.9 ± 0.8) × 10−4 L⊙, and the total dust
mass ranges from 5.7 × 1026 g for compact grains to 1.2 × 1027 g for highly porous grains, or ∼ 8 – 16 lunar masses.
Uncertainties in the dust luminosity and dust mass were derived by varying the model parameters and examining the
change in SED within flux uncertainties (10% at most wavelengths).
Dust Production Rates in the Fomalhaut Debris Disk
5
We also fit the observed SED (Figure 7) with astrosilicate grains using the Debris Disk Simulator (Wolf & Hillenbrand
2005) for two dust population corresponding to the size ranges a ≤ ablow and a ≥ ablow. The results from this modeling
are listed in (Table 2). The dust mass in blowout grains can be considered an upper limit because the SED can be
fit with a single power law size distribution with amin = ablow.The disk mass resulting from an astrosilicate model is
(6.6 ± 0.7) × 1025 g, or about 0.9 lunar masses. Although the estimated dust mass from the astrosilicate grain model
is lower than that for icy aggregate grains, the two compositions have similar masses for grains with 1 µm ≤ a ≤ 1
mm (Table 2).
5. DISK MODELING
5.1. Grain orbits
∗/r2R F (λ)πa2(Q(λ, a) + hcos θiQsca(λ, a)dλ (Tazaki & Nomura 2015). The scattering efficiencies Qsca(λ, a) were
In order to compute the timescales for mass loss in the disk from radiation pressure on dust grains, we consider
dust grains in orbit under the influence of gravity, radiation pressure, and centrifugal forces. Poynting-Robertson
drag at the cold belt is negligible. We used the equations of motion in Harwit (1988), modified by the term for
radiation pressure, to compute the orbits of blowout grains as conic sections from the cold belt (140 AU) to the
distant edge of the halo (450 AU). The incident radiation that contributes to radiation pressure on a grain is LRP =
R2
calculated from the analytical formulae for scattering of fluffy aggregates from Kataoka et al. (2014). Under the
assumption of intial Keplerian motion, the time was integrated numerically to find the blowout radius and blowout
time ∆t(a) (defined as the time for the grain to reach a distance 450 AU from the star) for each value of f . For
a hypothetical sample of small grains (a ≪ ablow), our calculations agree with an analytical calculation for the
blowout timescale from Lebreton et al. (2013), which is valid for cases where the radiation pressure dominates over
the gravitational force and Poynting-Robertson drag. We note our calculated value of 15 µm for the blowout size for
25% porosity differs slightly from that of Acke et al. (2012) (13 µm). This difference may be due to their inclusion of
iron sulfide in the grain composition. The presence of iron sulfide will increase the grain mass density and result in a
smaller blowout size. For blowout grains, ∆t(ablow) . 1100 yr, depending on grain size. For our minimum grain size
of 0.1 µm, ∆t ≈ 230 yr for f = 1.0 and ∆t ≈ 700 yr for f = 0.05.
The disk mass loss rate due to radiative blowout of small grains can be given as:
5.2. The Dust Production Rate
M =
a=ablowout
Xa=amin
N (a)m(a)/∆t(a)
(2)
Although larger grains are produced in the collisional cascade, they are removed more slowly than the smaller grains,
M for each value of f . For
and consequently they contribute far less to the mass loss rate. Table 2 gives the value of
M can be expressed as (1.2 ± 0.5) × 1021 g yr−1, with uncertainties derived from uncertainties in
fluffy aggregates,
porosity and observed fluxes.
For a disk containing an ongoing collisional cascade, we can compare the observed mass loss rate to a theoretical
dust production rate. In order to make this comparison, we scaled the range of predicted dust production rates for
0.01 – 1.0 µm grains from Kenyon & Bromley (2008) to our range of grain sizes with amin ≤ a ≤ ablow based on the
observed grain size distribution. The result yields the predicted dust mass that is produced every year for grain sizes
in the range amin ≤ a ≤ ablow. There is a range of predicted dust production rates to account for variations in initial
disk mass. The range of dust production rates corresponding to the observed grain size distribution for each value of
f are listed in Table 2. For icy grains, the predicted dust production rates agree with the observed mass loss rate. For
astrosilicates, the mass loss rate falls below the predicted range of dust production rates.
We can also compare the observed disk mass to the predicted disk mass from a collisional cascade (Kenyon & Bromley
2008). Table 2 lists the mass contained in compact grains with 1 µm ≤ a ≤ 1 mm and the mass in the corresponding
range for porous grains. The mass in these grains is 0.83 – 0.94 lunar masses. This range is a factor of ∼ 2 less than
the predicted mass of ∼ 2 lunar masses for a 1.7 M⊙ star (Kenyon & Bromley 2008).
6. DISCUSSION
6.1. The Nature of the Central Source
The nature of the inner region has been discussed by Absil et al. (2009), Acke et al. (2012), Lebreton et al. (2013),
and Su et al. (2013, 2016). Acke et al. (2012) used near-IR (AKARI and Absil et al. 2009) and Herschel data to
6
Adams et al.
measure the slope of the IR excess from ∼ 1 to 70 µm. They measured it as a power law in Fν with index -0.8. They
proposed this power law index was consistent with a stellar wind origin for the IR excess (Panagia & Felli 1975). If the
IR excess originates in a stellar wind, the power law predicts an IR excess of ∼ 0.2 – 0.3 Jy at 37 µm. Near-IR excesses
and the Herschel 70 µm excess flux from the central source source can be interpreted as originating from free-free
emission associated with a stellar wind. However, ALMA measurements of the central source at 870 µm (Su et al.
2016) indicate that the submillimeter flux is consistent with emission from a photosphere and not free-free emission.
In Figure 8, we show the mid-IR SED of the central source with photometry from SOFIA/FORCAST (this work) and
the predicted SED for free-free emission (Acke et al. 2012). The SOFIA/FORCAST photometry further rules out a
stellar wind origin for the IR excess.
Lebreton et al. (2013) proposed that the IR excess in the inner region originates from warm dust near the star at
1 – 2 AU. Such dust is within the ice sublimation radius. They predict an IR excess of ∼ 0.73 Jy at 37 µm, which
is slightly lower than the observed value, but still lies within the SOFIA/FORCAST uncertainties. However, these
models overpredict the flux at 10 – 24 µm (Su et al. 2016).
Su et al. (2016) also propose that the IR excess originates from warm dust, but they consider an asteroid belt at
the ice line (8 – 15 AU). Astronomical silicate dust is produced in this belt, with larger grains dragged inwards under
Poynting-Robertson drag. The dragged-in grains form a warm disk and might be halted in a pileup ring at 0.23 AU
due to density enhancement near the silicate sublimation radius. The presence of a pileup ring can cause additional IR
excess by ∼ 0.03 Jy over the ice-line dust and warm disk. Figure 8 also shows the SED for warm dust from Su et al.
(2013, 2016). The SOFIA/FORCAST data support the warm disk hypothesis. However, the SOFIA data do not
confirm or rule out the presence of a pileup ring, which results in additional excess of only 30 mJy at 37.1 µm.
6.2. Dust Production in the Kuiper Belt Analog
Acke et al. (2012) derived a mass loss rate of 2 × 1021 g yr−1 using fluffy aggregate grains with 25% porosity under a
M = 1 × 1021 g yr−1. We attribute
q = −3.5 size distribution. For 25% porosity, we derive the slightly lower value of
the difference in size distributions to additional constraints on the observed SED from Spitzer/MIPS (Stapelfeldt et al.
2004), SOFIA (this work), and ALMA (MacGregor et al. 2017).
Highly porous (91 – 94%) grains are observed in AU Mic (Graham et al. 2007) in scattered light. For Fomalhaut,
highly porous grains would imply that the size distribution extends to meter-sized objects, if the material is composed
of icy aggregates. Similar to mm-sized compact grains, such large objects are optically thick and highly (> 99%)
absorbing of radiation. For highly porous grains, the blowout size is & 100 µm and the minimum grain size is
∼ 60 µm.
in the Fomalhaut Kuiper Belt analog. Such a size distribution is consistent with the proposed belt of
micro-asteroids (Min et al. 2010).
If the grains are icy aggregates, the difference between the observed dust mass of ∼ 1 lunar mass and the value of
∼ 2 lunar masses predicted by Kenyon & Bromley (2008) is not unexpected. The predicted dust mass corresponds
to a different grain composition and porosity, with different absorption and scattering properties than the icy grains.
Consequently, the predicted mass loss rate and accumulated disk mass will differ from the observed values for icy
grains.
Icy aggregate dust grains can simultaneously reproduce the observed scattering and long wavelength thermal emis-
sion. If we extrapolate the predicted dust production rate for fine grains to our observed size distribution, we find
it to be consistent with the theoretical predictions of icy planet formation Kenyon & Bromley (2008). Therefore, the
icy grain hypothesis is consistent with the observed emission and the expected dust production rates. However, we
caution that the models of Kenyon & Bromley (2008) differ from the Fomalhaut debris disk. The Fomalhaut Kuiper
belt analog is narrower (13.5 ± 1.8 AU, MacGregor et al. 2017) than the theoretical disks, which span 30 – 150 AU,
so the theoretical disks have a larger volume of planetisimals that can be stirred by icy planets.
In addition, the
extrapolation of the dust production rates from fine (0.01 – 1.0 µm) grains to larger grains (a ≥ 0.1 µm) relies on
the assumption that dust is produced in steady state. An alternate interpretation could be that dust was produced
230 – 700 yr ago and grains with a < 0.1 µm have all been blown out of the system. In that case, the predicted dust
production rate cannot be extrapolated because the dust production is instantaneous, rather than steady state. The
steady state assumption implies that the minimum grain sizes are larger than ∼ 1 µm, or in the case of highly porous
grains, larger than ∼ 50 µm.
Both icy grain models and astrosilicate models can reproduce the observed SED. However, they differ in that the
astrosilicate model indicates a dearth of blowout grains. This result confirms the results of Pawellek et al. (2014). If
the grains are astrosilicates, they are produced at a lower rate or form with a minimum size that is larger than the
blowout size.
Dust Production Rates in the Fomalhaut Debris Disk
7
7. CONCLUSIONS
We have used SOFIA/FORCAST to observe Fomalhaut at 37.1 µm. The observations have constrained the nature
of the central source and the cold debris disk. The infrared excess of the point source at 37.1 µm is consistent with
the presence of an asteroid belt at the ice-line proposed by Su et al. (2013, 2016). The 37.1 excess in the point source
rules out a stellar wind as the origin of infrared excess in the central source.
Under the assumptions that the cold Fomalhaut debris disk is composed of icy aggregate dust grains and the disk
is in a steady state collisional cascade, our observed mass loss rate is consistent with theoretical predictions for dust
production rate from icy planet formation (Kenyon & Bromley 2008). Icy grains remain a plausible dust grain type for
the Fomalhaut because their presence around Fomalhaut would explain the observed scattered light, thermal emission,
and the observed dust production rate. However, if the dust is composed of astrosilicates, the observed mass loss rate
is less than the predicted range of dust production rates.
We thank K. Stapelfeldt for sharing his PSF-subtracted MIPS image of Fomalhaut and K. Su for sharing her inner disk
model SEDs. We thank the SOFIA ground crew, flight crew, and Mission Operations for their successful execution
of the SOFIA observations. We also thank an anonymous referee whose suggestions led to the improvement of this
manuscript. This work is based on observations made with the NASA/DLR Stratospheric Observatory for Infrared
Astronomy (SOFIA). SOFIA science mission operations are conducted jointly by the Universities Space Research
Association, Inc. (USRA), under NASA contract NAS2-97001, and the Deutsches SOFIA Institut (DSI) under DLR
contract 50 OK 0901.
Facility: Facilities: Spitzer, SOFIA, Herschel
8
Adams et al.
REFERENCES
Absil, O., Mennesson, B., Le Bouquin, J.-B. et al. 2009, ApJ,
704, 150
Acke, B., Min, M., Dominik, C. et al. 2012, A&A, 540, 125
Aumann, H. H. 1985, PASP, 97, 885
Castelli, F. & Kurucz, R. L. 2004, arXiv:0405087
Dohnanyi, J. S. 1969, J. Geophys. Res., 74, 2531
Draine, B. T. 2003, ARA&A, 41, 241
Fraundorf, P., Walker, R. M., & Brownlee, D. E. 1982, in Comet
Discoveries, Statistics, and Observational Selection, ed. L. L.
Wilkening, IAU Colloq. 61, 383
MacGregor, M. A., Matr`a, L., Kalas, P. et al. 2017, ApJ, 842, 8
Mamajek, E. E. 2012, ApJL, 754, L20
Marsh, K. A., Velusamy, T., Dowell, C. D. et al. 2005, ApJL,
620, L47
Min, M., Kama, M., Dominik, C., & Waters, L. B. F. M. 2010,
A&A, 509, L6
Panagia, N. & Felli, M. 1975, A&A, 39, 1
Pawellek, N., Krivov, A., Marshall, J. et al. 2014, ApJ, 792, 65
Poglitsch, A., Waelkens, C., Geis, N. et al. 2010, A&A, 518, L2
Pollack, J. B., Hollenback, D., Beckwith, S. et al. 1994, ApJ,
Graham, J. R., Kalas, P. G. & Matthews, B. C. 2007, ApJ, 654,
421, 615
595
Griffin, M. J., Abergel, A., Abreu, A. et al. 2010, A&A, 518L, 3
Harwit, M. 1998, Astrophysical Concepts (2nd ed.; New York:
Springer-Verlag)
Herter, T. L., Adams, J. D., De Buizer, J. M. et al. 2012, ApJL,
749, L18
Stapelfeldt, K. R., Holmes, E. K., Chen, C. et al. 2004, ApJS,
154, 458
Su, K. Y. L., Rieke, G. H., Malhotra, R. et al. 2013, ApJ, 763,
118
Su, K. Y. L., Rieke, G. H., Defr´ere, D. et al. 2016, ApJ, 818, 45
Su, K. Y. L., De Buizer, J. M., Rieke, G. H. et al. 2017, AJ, 153,
Herter, T. L., Vacca, W. D., Adams, J. D. et al. 2013, PASP,
226
125, 1393
Holland, W. S., Greaves, J. S., Dent, W. R. F. et al. 2003, ApJ,
582, 1141
Kalas, P., Graham, J. R., & Clampin, M. 2005, Nature, 435, 1067
Kalas, P., Graham, J. R., Fitzgerald, M. P., & Clampin, M.
2013, ApJ, 775, 56
Kataoka, A., Okuzumi, S., Tanaka, H. et al. 2014, A&A, 568, 42
Kenyon, S. J. & Bromley, B. C. 2008, ApJS, 179, 451
Lau, R. M., Herter, T. L., Morris, M. R. et al. 2013, ApJ, 775, 37
Lebreton, J., van Lieshout, R., Augereau, J.-C. et al. 2013, A&A,
555, A146
Tazaki, R. & Nomura, N. 2015, ApJ, 799, 119
van Leeuwen, F. 2007, A&A, 474, 653
Wolf, S. & Hillenbrand, L. A. 2005, Comp. Phys. Comm., 171,
208
Young, E. T., Becklin, E. E., De Buizer, J. M. et al. 2012, ApJL,
749, L17
Dust Production Rates in the Fomalhaut Debris Disk
9
Figure 1. Left: Processed image of Fomalhaut at 37.1 µm. The dashed green ellipse is the approximate location of the 70 µm
cold belt from (Acke et al. 2012). Right: Smoothed image of Fomalhaut at 37.1 µm with the same cold belt location. The
smoothing kernel was a Gaussian with FWHM of 6.2′′ × 6.2′′.
10
Adams et al.
Figure 2. PSF fitting and subtraction for α Boo. Upper left: FORCAST image of α Boo at 37.1 µm, averaged over the four
OC4-G flights. Lower left: Radial plot for α Boo (black dots), with flux densities in instrumental units. Model fits to the data
(red points) include contributions from elliptical Gaussian jitter (green points) and a modified Airy function for diffraction and
non-Gaussian components (blue points). Upper right: Image of α Boo at 37.1 µm after PSF subtraction. Lower right: Radial
plot of residuals after PSF subtraction.
Dust Production Rates in the Fomalhaut Debris Disk
11
Figure 3. Analogous to Figure 2, but for Fomalhaut. The jitter model parameters were adjusted to fit the Fomalhaut data.
The Fomalhaut PSF model uses the same modified Airy funtion as the α Boo PSF model.
12
Adams et al.
Figure 4.
Integrated brightness profile along the apparent major axis, where the positive direction corresponds to the SE
direction, and the negative direction corresponds to the NW direction. The width for the integration was 23.′′9 (31 pixels).
Profiles with (dotted line) and without (solid line) the PSF subtracted are shown. The edges of the ansai at 17′′ and −19.3′′are
noted with dashed lines.
Dust Production Rates in the Fomalhaut Debris Disk
13
Figure 5. Grain equilibrium temperatures T (a) vs. grain radius a for icy grains with volume filling factors f = 1.0, 0.5, 0.1,
and 0.05 at a distance of 140 AU.
14
Adams et al.
Figure 6. Observed SED and best-fitting models for the icy aggregate grain compositions. The SEDs for icy grains over a range
of porosities (f = 1.0, 0.5, 0.1, and 0.05) (Kataoka et al. 2014) are shown. The contributions from blowout grains (dotted lines)
and large grains (dashed lines) are shown, with the total SED depicted as solid black lines. Blue bar: Spitzer/MIPS flux from
photosphere-subtracted image (Stapelfeldt et al. 2004). Green bar: FORCAST data (this work). Red bars: Herschel model for
the outer disk from Acke et al. (2012). Cyan bar: ALMA data at 1.3 mm from MacGregor et al. (2017).The submillimeter data
are at 350 µm (Marsh et al. 2005), and 450 and 850 µm (Holland et al. 2003). The submillimeter fluxes are shown as upper
limits because they include contamination from the inner disk.
Dust Production Rates in the Fomalhaut Debris Disk
15
Figure 7. Observed SED and model fits for astrosilicates. Lines and data points are the same as those in Figure 6.
16
Adams et al.
Figure 8. Infrared excess vs. wavelength for the point source at 24 µm (blue point, Su et al. 2016), 37.1 µm (green point, this
work), and 70 µm (red point, Acke et al. 2012, Su et al. 2013). The total emission from the inner disk model of Su et al. (2016)
is shown both with (dot-dashed line) and without (dotted line) the pileup ring. The expected infrared excess from a stellar wind
(dahsed line) (Acke et al. 2012) is shown.
Dust Production Rates in the Fomalhaut Debris Disk
17
Table 1. Summary of SOFIA/FORCAST observations of Fomalhaut under Plan ID 04 0064 and AOR
ID 04 0064 1.
Mission ID
Start time (UT)
End time (UT)
Altitude (ft)
Starting Elevation angle (◦ )
Calibration uncertainty (%)
2016-07-12 FO F319
2016-07-12 FO F319
09:50:35
10:10:05
2016-07-12 FO F319
10:54:56
2016-07-14 FO F321
2016-07-14 FO F321
10:49:47
11:22:55
2016-07-14 FO F321
11:48:33
2016-07-14 FO F321
11:48:33
2016-07-14 FO F321
2016-07-14 FO F321
12:44:27
13:31:10
2016-07-14 FO F321
13:58:29
2016-07-18 FO F323
2016-07-18 FO F323
10:56:20
11:29:10
2016-07-18 FO F323
11:59:54
2016-07-18 FO F323
2016-07-18 FO F323
12:25:43
12:51:40
2016-07-20 FO F325
10:07:12
2016-07-20 FO F325
10:46:56
2016-07-20 FO F325
2016-07-20 FO F325
11:26:55
12:06:35
2016-07-20 FO F325
12:41:21
10:10:05
10:54:56
11:44:03
11:22:55
11:48:33
12:16:41
12:44:16
13:31:10
13:44:20
14:10:08
11:29:10
11:59:41
12:25:43
12:51:40
13:15:49
10:46:56
11:26:44
12:04:21
12:39:23
13:16:46
41000
42000
41000
38000
39000
39000
40000
41000
42000
42000
40000
40000
41000
42000
42000
39000
39000
39000
39000
41000
30
33
39
26
33
39
46
50
57
59
25
31
38
43
49
34
41
48
52
55
6
6
6
5
5
5
5
5
6
6
5
5
5
5
5
5
5
5
5
5
1
8
A
d
a
m
s
e
t
a
l
.
M is the mass loss rate due to radiation pressure,
Table 2. Summary of dust model parameters and results for the cold region. Md is the dust mass,
MDP is the scaled dust production rate for grains with amin ≤ a ≤ ablow from Kenyon & Bromley (2008), and M1mm is the mass in 1 – 1000 µm (scaled
with porosity) grains.
Large grains
Blowout grains
Composition
f
icy aggregates
icy aggregates
icy aggregates
1.0
0.5
0.1
icy aggregates 0.05
ablow
(µm)
10
21
100
210
astrosilicates
1.0
5
q
amin
amax
log Md
q
amin
amax
log Md
(µm)
(m)
(g)
(µm)
(µm)
(g)
-3.4
-3.4
-3.4
-3.4
-3.6
12
40
280
540
5.0
0.1
0.7
0.04
26.78+0.14
−0.02
26.97+0.04
−0.03
27.07+0.02
−0.03
27.03+0.02
1.2
−0.04
0.001 25.82+0.03
−0.04
-3.5
-3.5
-3.5
-3.5
-3.6
0.1
0.1
0.1
0.1
0.01
12
40
280
540
5.0
23.62+0.20
−0.02
23.78+0.28
−0.20
24.02+0.05
−0.11
23.99+0.06
−0.18
log M
(g yr−1)
20.96+0.02
−0.11
21.00+0.03
−0.19
21.17+0.04
−0.11
21.15+0.07
−0.18
log MDP
(g yr−1)
M1mm
(g)
[20.65, 21.11] 25.77+0.03
−0.03
[20.66, 21.12] 25.83+0.04
−0.03
[20.68, 21.14] 25.82+0.03
−0.03
[20.69, 21.15] 25.84+0.02
−0.04
[20.14, 20.60] 25.81+0.03
−0.04
≤ 21.86
≤ 19.38
|
1801.10079 | 1 | 1801 | 2018-01-30T16:12:15 | High Abundances of Presolar Grains and $^{15}$N-rich Organic Matter in CO3.0 Chondrite Dominion Range 08006 | [
"astro-ph.EP"
] | NanoSIMS C-, N-, and O-isotopic mapping of matrix in CO3.0 chondrite Dominion Range (DOM) 08006 revealed it to have in its matrix the highest abundance of presolar O-rich grains (257 +76 / -96 ppm, 2$\sigma$) of any meteorite. It also has a matrix abundance of presolar SiC of 35 (+25 / -17, 2$\sigma$) ppm, similar to that seen across primitive chondrite classes. This provides additional support to bulk isotopic and petrologic evidence that DOM 08006 is the most primitive known CO meteorite. Transmission electron microscopy of five presolar silicate grains revealed one to have a composite mineralogy similar to larger amoeboid olivine aggregates and consistent with equilibrium condensation, two non-stoichiometric amorphous grains and two olivine grains, though one is identified as such solely based on its composition. We also found insoluble organic matter (IOM) to be present primarily as sub-micron inclusions with ranges of C- and N-isotopic anomalies similar to those seen in primitive CR chondrites and interplanetary dust particles. In contrast to other primitive extraterrestrial materials, H isotopic imaging showed normal and homogeneous D/H. Most likely, DOM 08006 and other CO chondrites accreted a similar complement of primitive and isotopically anomalous organic matter to that found in other chondrite classes and IDPs, but the very limited amount of thermal metamorphism experienced by DOM 08006 has caused loss of D-rich organic moieties, while not substantially affecting either the molecular carriers of C and N anomalies or most inorganic phases in the meteorite. One C-rich grain that was highly depleted in $^{13}$C and $^{15}$N was identified; we propose it originated in the Sun's parental molecular cloud. | astro-ph.EP | astro-ph | High Abundances of Presolar Grains and 15N-rich Organic Matter in CO3.0 Chondrite
Dominion Range 08006
Larry R. Nittler1, Conel M. O'D. Alexander1, Jemma Davidson1, My E. I. Riebe1, Rhonda M.
Stroud2, and Jianhua Wang1
1Department of Terrestrial Magnetism, Carnegie Institution of Washington, Washington, DC
20015, USA.
2Materials Science and Technology Division, Code 6366, US Naval Research Laboratory,
Washington, DC 20375-5320, USA.
To appear in Geochimica Cosmochimica Acta, submitted Oct 31, 2017; accepted January 30, 2018
Abstract
NanoSIMS C-, N-, and O-isotopic mapping of matrix in CO3.0 chondrite Dominion Range
(DOM) 08006 revealed it to have in its matrix the highest abundance of presolar O-rich grains
(257 +76/-96 ppm, 2σ) of any meteorite. It also has a matrix abundance of presolar SiC of 35
(+25/-17, 2σ) ppm, similar to that seen across primitive chondrite classes. This provides additional
support to bulk isotopic and petrologic evidence that DOM 08006 is the most primitive known CO
meteorite. Transmission electron microscopy of five presolar silicate grains revealed one to have
a composite mineralogy similar to larger amoeboid olivine aggregates and consistent with
equilibrium condensation, two non-stoichiometric amorphous grains and two olivine grains,
though one is identified as such solely based on its composition. We also found insoluble organic
matter (IOM) to be present primarily as sub-micron inclusions with ranges of C- and N-isotopic
anomalies similar to those seen in primitive CR chondrites and interplanetary dust particles. In
contrast to other primitive extraterrestrial materials, H isotopic imaging showed normal and
homogeneous D/H. Most likely, DOM 08006 and other CO chondrites accreted a similar
complement of primitive and isotopically anomalous organic matter to that found in other
chondrite classes and IDPs, but the very limited amount of thermal metamorphism experienced by
DOM 08006 has caused loss of D-rich organic moieties, while not substantially affecting either
the molecular carriers of C and N anomalies or most inorganic phases in the meteorite. One C-rich
1
grain that was highly depleted in 13C and 15N was identified; we propose it originated in the Sun's
parental molecular cloud.
1. INTRODUCTION
Chondritic meteorites and micrometeorites, stratospheric interplanetary dust particles (IDPs)
and dust from comet Wild 2 brought to Earth by NASA's Stardust spacecraft all contain traces of
materials that pre-date the formation of the Sun's protoplanetary disk. Presolar stardust grains
originated as condensates in the outflows and ejecta of evolved stars. They are recognized by their
highly anomalous isotopic compositions, which largely reflect nuclear processes that occurred in
the grains' parent stars (Zinner, 2014; Nittler and Ciesla, 2016). Additionally, a large fraction of
the C present in primitive meteoritic materials is in the form of a macromolecular organic matter
("insoluble organic matter" or IOM) that shows isotopic anomalies (typically high D/H and/or
15N/14N ratios relative to terrestrial values) strongly indicative of an origin in either the Sun's
parental molecular cloud, or in the outer reaches of the nascent Solar System, where conditions
were similar to those in molecular clouds (Messenger, 2000; Busemann et al., 2006).
A large number of presolar stardust phases have been identified, the most abundant being
crystalline and amorphous silicates, Al2O3, MgAl2O4, and SiC, but also including graphitic grains
(often containing sub-grains of metal and carbides), Si3N4, and other minor phases (Zinner, 2014).
Nanometer-sized diamonds (nanodiamonds) carrying isotopically anomalous noble gases are also
abundant, but their small sizes make their identification as presolar grains ambiguous (Dai et al.,
2002; Stroud et al., 2011; Heck et al., 2014). Presolar grains are essentially fossil remnants of stars
that survived interstellar and solar nebular processing, and as such they have been used as tools to
probe a wide variety of astrophysical processes. Because they encompass such a wide range of
chemical forms, they also can serve as sensitive probes of processes that affected materials in the
interstellar medium (e.g., shocks and irradiation, Vollmer et al., 2007), the protoplanetary disk
(e.g., heating, Huss and Lewis, 1995), and in asteroidal and cometary parent bodies (e.g., aqueous
alteration, thermal metamorphism: Floss and Stadermann, 2012; Leitner et al., 2012b). In
particular, detailed analyses of presolar grain abundances in meteorites suggest that the different
chondrite groups accreted a similar mix of presolar materials and abundance variations observed
2
today reflect that different phases respond differently to parent-body processes. For example, the
low-temperature aqueous alteration that has affected many carbonaceous chondrites has little or
no obvious impact on SiC abundances (Davidson et al., 2014a), but can be very destructive of
presolar silicates (Leitner et al., 2012b). Thermal metamorphism destroys all presolar grains, but
at different rates. For instance, presolar MgAl2O4 is in much higher abundance in meteorites that
have seen lower degrees of heating, like CM chondrites (Zinner et al., 2003) and the LL3 Krymka
(Nittler et al., 2008), than in the more heated Tieschitz (H/L3.6) ordinary chondrite (Nittler et al.,
1997). Of the known stardust phases, silicates are the most susceptible to destruction by parent-
body processing and their abundances (and the ratio of their abundances to those of the more
resistant oxide phases) are thus highly sensitive indicators of the degree to which the material they
are embedded in has been processed.
The highest presolar silicate abundances have been observed in anhydrous fine-grained IDPs
collected in Earth's stratosphere and micrometeorites collected in Antarctica, with reported
abundances ranging from a few hundred ppm to percent levels (e.g., Floss et al., 2006; Busemann
et al., 2009; Davidson et al., 2012; Alexander et al., 2017c). Floss et al. (2006) calculated an
average presolar silicate abundance of 375 ppm for a sub-set of 15N-rich IDPs, termed "isotopically
primitive." Anhydrous IDPs are suspected to have originated from comets and the high abundance
of presolar grains supports this hypothesis, as comets are expected to be more primitive than the
asteroidal parent bodies of chondritic meteorites. Prior to the present study, the highest presolar
silicate abundances in meteorites have been reported for the highly primitive ungrouped
carbonaceous chondrite Acfer 094 (150-200 ppm; Nguyen et al., 2007; Vollmer et al., 2009b;
Hoppe et al., 2015), the CO 3.0 chondrite Allan Hills (ALH) 77307 (190 ppm; Nguyen et al.,
2010), and the relatively unaltered CR2 chondrites Queen Alexandra Range (QUE) 99177 and
Meteorite Hills (MET) 00426 (160-220 ppm; Floss and Stadermann, 2009a; Nguyen et al., 2010).
Because presolar grains reside in the fine-grained matrix between chondrules and other inclusions
in chondrites, these quoted abundances are all matrix-normalized, but are still lower than the
average abundance reported for IDPs. These lower abundances indicate that there has been
significant destruction of presolar silicates either in the nebular regions where the chondrites
accreted or during the very minor degrees of alteration that petrographic studies indicate occurred
in the parent bodies of these four meteorites (Brearley, 1993; Greshake, 1997; Harju et al., 2014).
3
An additional type of primitive, possibly presolar, material found in meteorites and IDPs is
isotopically anomalous organic matter (Messenger, 2000; Busemann et al., 2006; Alexander et al.,
2007; Davidson et al., 2012; Alexander et al., 2017). Macromolecular insoluble organic matter
(IOM) is the dominant form of C in primitive chondrites and IDPs. Large isotopic variations, most
notably enrichments in D and 15N, relative to terrestrial D/H and 15N/14N ratios, are commonly
observed in the bulk isotope compositions for organic matter from different parent bodies, and
even greater variations often occur at the µm to sub-µm scale in the most primitive samples
(Messenger, 2000; Keller et al., 2004; Busemann et al., 2006; Alexander et al., 2007; Davidson et
al., 2012; De Gregorio et al., 2013). Carbon-isotopic anomalies are also seen occasionally in IOM
from IDPs and primitive meteorites (Floss et al., 2004; Busemann et al., 2006; Floss and
Stadermann, 2009b). In situ studies (e.g., Busemann et al., 2006; Remusat et al., 2010; Bose et al.,
2014; Le Guillou and Brearley, 2014) of chondrites indicate that IOM is generally present as
discrete, typically sub-µm, grains, including "nanoglobules," which are spherical and often hollow
solid organic grains (Nakamura-Messenger et al., 2006; De Gregorio et al., 2013). How the
meteoritic IOM formed is still an open question. The large D and 15N enrichments as well as
infrared spectral similarities with dust in the diffuse interstellar medium have long been interpreted
to indicate an interstellar heritage for the IOM (Robert and Epstein, 1982; Yang and Epstein, 1983;
Pendleton et al., 1994), or that it formed in parent bodies from originally interstellar material (Cody
et al., 2011; Vollmer et al., 2014). Alternatively, it has been suggested that the IOM may have
formed in the solar nebula, e.g., by irradiation of simpler molecular ice precursors or by Fischer-
Tropsch type processes (Gourier et al., 2008; Nuth et al., 2008; Ciesla and Sandford, 2012).
Complementary to presolar grain abundances, the nature of IOM, e.g. abundance, isotopic
composition, and chemical structure, can also be a sensitive indicator of thermal metamorphism
and aqueous alteration in meteorite parent asteroids (Quirico et al., 2003; Alexander et al., 2007;
Busemann et al., 2007; Cody et al., 2008; Alexander et al., 2013; Bonal et al., 2016; Alexander et
al., 2018).
In this paper, we report the discovery that the Antarctic CO3 chondrite Dominion Range
(DOM) 08006 has a higher matrix-normalized abundance of O-rich presolar stardust grains than
any previously studied chondrite. Moreover, we found DOM 08006 to contain abundant organic
material exhibiting a similar range of C- and N-isotopic anomalies to that seen in primitive CR
and CM chondrites (albeit with much more muted D anomalies than seen in those meteorite
4
classes). That DOM 08006 may be special was indicated by a prior study of amino acids in a
number of CO chondrites by Burton et al. (2012). These authors found that it contained a lower
abundance of straight-chain amino acids (associated with thermal alteration) than ALH 77307,
suggesting that DOM 08006 may have experienced less heating than ALH 77307, hitherto
considered the least altered of the CO3 chondrites (Grossman and Brearley, 2005). Spurred on by
this work, Alexander et al. (2018) included DOM 08006 in a survey of bulk isotopic and
mineralogical properties of primitive CO chondrites and found that this meteorite has a
significantly higher C content and slightly higher H/C, D, and 15N contents compared to ALH
77307, indicating that it has preserved more primitive organic matter. The high abundance of
presolar materials reported here, together with detailed bulk chemical and isotopic, mineralogical
and petrologic data for this meteorite reported by Davidson et al. (2014b) and Alexander et al.
(2018), bears this out and indicates that DOM 08006 is among the least altered of any chondritic
meteorite. It is thus a very valuable specimen both for understanding the original unaltered nature
of nebular materials that accreted onto planetesimals in the protoplanetary disk as well as the
earliest stages of alteration on asteroids.
2. SAMPLE AND EXPERIMENTAL METHODS
We obtained a polished thin section (DOM 08006, 16) from the Johnson Space Center (Figure
1) and analyzed it by isotopic imaging with the Cameca NanoSIMS 50L ion microprobe at the
Carnegie Institution of Washington. We used very similar methods to those used previously for
presolar grain and organic mapping of meteorite thin sections (Nguyen et al., 2010). For most
measurements, we rastered a focused Cs+ primary ion beam over 10µm×10µm areas with
simultaneous collection of negative secondary ions of the C isotopes (either 12,13C or 12C2, 12C13C),
16,17,18O, 28Si, either 12C14N or 27Al16O on individual electron multipliers (EMs) along with
secondary electrons. The effective spatial resolution (taking into account beam size and slight
broadening due to inaccuracy in correcting for stage/beam shifts over the long measurements) for
most images was between 100 and 150 nm. One set of images was acquired in 16,17O, 12C2, 12C13C,
12C14N, 12C15N, 28Si negative ions and secondary electrons to characterize the N isotopic
composition of carbonaceous material. For all measurements, the mass resolving power was set so
that isobaric interferences (e.g., 16OH on 17O, 13C2 on 12C14N) were negligible and the primary
beam current was set (~1.2 pA) such that the maximum 16O count rate in each image was typically
5
~106 counts per second. Twenty sequential 256×256 pixel images (cycles) were acquired on each
area, for a total acquisition time of 65 minutes (total dwell time of 60 milliseconds per 39 nm × 39
nm pixel). A Nuclear Magnetic Resonance probe was used to maintain magnetic field stability in
the mass spectrometer, and the peak positions of 12C and 16O were checked every five cycles and
the other peaks shifted accordingly. Prior to image acquisition, areas were pre-sputtered for
roughly three minutes with a more intense primary beam to remove the C coat and implant Cs,
which greatly boosts the yield of negative secondary ions.
In addition, two sets of imaging runs on the DOM 08006 thin section were carried out with a
higher (~10 pA) Cs+ beam current to analyze H isotopes together with those of C (1,2H, 12,13C, 16O,
and 12C14N). The spatial resolution was degraded to ~350 nm due to the higher primary beam
current, and slightly longer dwell times were also used (100 ms total counting time per pixel).
We used the L'IMAGE software (L. R. Nittler) to analyze the image data. Images were
corrected for the 44 ns deadtime of the EM counting system and shifts between frames were
corrected for via an autocorrelation algorithm. We also corrected the O-isotopic images for quasi-
simultaneous arrival (QSA) effects (Slodzian et al., 2004). QSA refers to the possibility that a
single primary ion may produce two or more secondary ions of a given species within a time
interval much shorter than the deadtime of the NanoSIMS EM counting system. Since the
probability of this occurring is much higher for 16O than for 17O or 18O, especially with the high
secondary ion transmission of the NanoSIMS, this can lead to undercounting of 16O and hence
introduce an isotopic fractionation. Slodzian et al. (2004) showed that, according to Poisson
statistics, the magnitude of the fractionation should be proportional to the ratio of secondary ion
to primary ion count rates, but empirically the proportionality constant does not in general match
the prediction. We applied the formalism of Slodzian et al. (2004) to 18O/16O ratio images to
empirically determine and correct 16O images for the appropriate QSA proportionality constant.
We generally calculated the constant from the individual pixels of the first image acquired in an
area of the sample and used it to correct all the images for that run. The maximum correction for
a given image pixel was typically less than 10%. Both the deadtime and QSA corrections were
carried out on a pixel-by-pixel basis on individual image frames.
Following the deadtime, QSA and alignment corrections, we calculated pixel-by-pixel isotopic
ratio images (e.g., 17O/16O, 18O/16O, 13C/12C, 15N/14N, D/H) after first applying a 3×3 pixel boxcar
6
smoothing to the images to improve the signal to noise ratio and to remove spurious variations
below the true spatial resolution. Oxygen isotopes were internally normalized to the average
composition of each image, H and C isotopes were normalized to epoxy infilling cracks in the thin
section. Comparison to measurements of IOM from the CR chondrite QUE 99177 used as an
external standard indicated that the epoxy has δD and δ13C within 10% and 5% of zero,
respectively. Nitrogen isotopes were normalized so that the average composition of carbonaceous
particles in an image matched the bulk value of δ15N=10.5 ‰ for IOM from this meteorite
(Alexander et al., 2018). Presolar O-rich and C-rich grain candidates were identified by manual
examination of the isotopic ratio images and of "sigma" images, in which each pixel represents
the number of standard deviations (based on Poisson statistics) its measured isotopic ratio is away
from terrestrial values (Figure 2). Candidate grain regions of interest (ROIs) were selected as
regions of several contiguous pixels clearly deviating from the standard ratio(s). In general, only
pixels within the full-width at half maximum of the anomalous region in a "sigma" image were
included in an ROI (Fig. 2). Isotopic ratios were then calculated from the summed counts of all
pixels within an ROI, which was confirmed to be a presolar grain candidate if at least one isotopic
ratio was >3σ away from the surrounding meteorite matrix. Some candidates, including all grains
with anomalous 17O/16O identified during the O-N runs as well as several grains for which the
apparent anomaly was marginal (e.g., <4σ), were subsequently re-measured for all three O
isotopes, again in imaging mode but with a smaller raster, typically 3µm×3µm. To characterize
organic materials, we used an automatic particle definition algorithm (Nittler et al., 1997) to define
individual C-rich ROIs in the images and calculated C-, N- and/or H-isotopic ratios from the
summed pixel counts. This was only done for the combined C and N and combined H and C runs.
Note that because the NanoSIMS primary ion beam tails will always contribute some signal from
surrounding material to the measurement of grains of interest, all reported isotopic anomalies are
lower limits.
Following the NanoSIMS measurements, the imaged areas were examined in a JEOL 6500F
scanning electron microscope (SEM) equipped with an energy dispersive X-ray analysis system.
Four cross-sections sampling five presolar silicate grains were extracted by focused ion beam
lift-out with a FEI Nova 600 FIB-SEM at the Naval Research Laboratory (NRL) for analysis by
transmission electron microscopy (TEM). Analytical TEM studies were performed with the JEOL
7
2200FS field-emission scanning transmission electron microscope at NRL, equipped with a Noran
System Six energy dispersive X-ray spectrometer. EDS spectra of individual grains were
quantified with Cliff-Lorimer routines, with K factors calibrated from San Carlos olivine and
Tanzanian hibonite standards.
3. RESULTS
3.1. NanoSIMS imaging
We mapped four areas of DOM 08006 fine-grained matrix that were separated from one
another by several mm (Fig. 1) in case of heterogeneity across the section. A total of ~28,000 µm2
was covered for O and C isotopes, including about 700 µm2 measured for N isotopes as well
(Section 2). A total of 6200 µm2 were covered for H isotopes. Some details of the mapping results
are provided in Table 1. We note that following the acquisition of O and C data for Area 7, we
found that the primary beam had inadvertently been slightly defocused so that the lateral resolution
for these images was ~300 nm, compared to the resolution of 100-150 nm for the other runs.
Example images of one particularly interesting 10 × 10 µm2 region in Area 3 containing a presolar
silicate grain, a presolar SiC grain and isotopically anomalous organic material, are shown in
Figure 3.
3.1.1. O-rich presolar grains
Analysis of the NanoSIMS images revealed a total of 101 grains with substantial anomalies
(>4 σ) in their O-isotopic ratios indicating that they are presolar, circumstellar grains. Data for
these are summarized in Table 2. Grain diameters range from 150 nm up to 1 µm (Figure 4). For
most grains, these were estimated from the NanoSIMS images. If N pixels were included in a grain
ROI, the equivalent diameter is calculated from 2×√(𝑁𝑁×𝑎𝑎𝜋𝜋), where a is the area of a single
image pixel, equal to 0.00152 µm2 for most of the images. However, the apparent size of
isotopically anomalous regions in NanoSIMS images depends both on the intrinsic size of the
anomalous grain and the size of the approximately Gaussian-shaped primary ion beam. We thus
report two sizes for the grains in Table 2: the original size inferred from the images and a corrected
8
size, where an assumed 120-nm beam diameter is subtracted in quadrature from the original
diameter. In almost all previous presolar grain studies of meteorites and IDPs, only the former
values are reported. The one exception is the study of Zhao et al. (2013), who applied an analogous
correction to grain sizes determined by NanoSIMS based on comparison with images from the
Auger microprobe, although their correction was apparently slightly larger, 20-30%, than ours.
The average measured grain diameter is 293 nm, whereas this is decreased in our study by ~10%
to 264 nm for the corrected data. We consider the corrected grain sizes to be conservative lower
limits.
With the exception of the five grains for which we have TEM data (Section 3.2), we have little
mineralogical information about the identified presolar O-rich grains. We can glean some
elemental information from the NanoSIMS data. For 65 of the presolar grains, the NanoSIMS
imaging runs included measurement of both 28Si- and 27Al16O- secondary ions. The measured Si-
/O- secondary ion ratios are plotted against the AlO-/O- ion ratios for these grains in Figure 5b. All
of the grains show Si/O ion ratios in the range we typically observe for silicates. However, since
the vast majority of known presolar oxide grains are Al-rich (Nittler et al., 2008), we assumed that
the grains with high AlO/O ratios are oxides and, based on a break in the histogram of AlO/O
values (Figure 5a), chose a threshold of AlO/O=0.007, above which a grain may be identified as
an oxide. Most of these grains have Si/O ratios at the lower end of the observed range, consistent
with their measured Si being contributed from surrounding material during the NanoSIMS imaging
measurements. Consequently, as an additional criteria for identifying oxide grains, we included
only those grains with the lowest Si/O ratios, i.e., below 0.01 (grey box in Figure 5). With this
threshold, 10 of the 65 grains are inferred to be oxides for an overall ratio of presolar silicates to
oxides of 5.5. The uncertainty on this number is obviously quite large, however. For example,
changing the Al/O threshold to 0.01 would raise the presolar silicate to presolar oxide ratio to ~12.
Moreover, Nguyen et al. (2010) found that 6 of 10 presolar grains identified as oxides from
NanoSIMS elemental images were subsequently revealed to be silicates by Auger microprobe
analysis, and our inferred silicate/oxide ratio is thus likely a lower limit.
The O-isotopic ratios for the 101 identified O-rich presolar grains are compared with previous
data for presolar oxides and silicates in Figures 6 and 7. It has long been recognized that the data
for presolar silicates show narrower ranges of O isotopic ratios compared to the presolar oxide
9
database (e.g., Nguyen et al., 2007). This is a direct result of the fact that the presolar silicates have
almost exclusively been found by in situ isotopic imaging methods, whereas a large number of the
oxides were measured as single isolated grains in meteoritic acid residues. Even for nominally
resolved grains, the tails of the NanoSIMS primary ion beam contribute some atoms from
isotopically normal surrounding material to the measurements of the presolar grains (Nguyen et
al., 2007), a phenomenon usually referred to as "isotopic dilution." This has a larger effect on
grains with 17O and/or 18O depletions, and the lack of presolar silicates with 18O/16O ratios less
than about 5×10-4, compared to the presolar oxides (Figures 6 and 7), is thus almost certainly an
effect of isotope dilution, not a sign that such silicate grains are not present in the meteorites.
The DOM 08006 presolar grains are in generally good agreement with the previous data, but
show slightly narrower ranges of O-isotope ratios compared to the larger presolar silicate database
(bottom panels of Figure 7). It is unlikely that this represents a true difference in the isotopic
compositions of presolar grains in this meteorite. More likely, it indicates both the more limited
statistics of the present study and that isotopic dilution played a larger role in our O-isotopic
measurements than in some previous studies. For all of the imaged areas except Area 7, our spatial
resolution of 100–150 nm was comparable to that used in most other NanoSIMS in situ studies
(but see Hoppe et al., 2015, 2017 for studies based on higher resolution). However, the measured
isotopic composition of a grain analyzed in the NanoSIMS imaging mode depends critically on
which pixels are included in the ROI calculation; including more pixels will generally increase
statistical precision but also enhance the effect of isotope dilution. Few published studies of
presolar silicates have specified what criteria were used to include pixels in the isotope ratio
calculations. We included all pixels within the full width at half maximum of "sigma" images
(Section 2, Fig. 2) and for many of the grains, the derived ratios would have been more anomalous
had we chosen to restrict the calculation to fewer pixels. Therefore, it is likely that the narrower
span of O isotopic ratios measured in DOM 08006 grains compared to the prior presolar silicate
database is a result of our approach to defining grain ROIs. However, we note that essentially all
but the largest grains are compromised by isotope dilution in all in situ studies (Nguyen et al.,
2007; Nguyen et al., 2010), regardless of image processing choices, and almost all of the grains
thus have true O-isotope anomalies that are more extreme than reported.
10
3.1.2. C- and N-isotopic mapping
We identified a total of 66 sub-µm carbonaceous inclusions with anomalous 13C/12C and/or
15N/14N ratios (Table 3). An ROI was identified as anomalous if its C isotopic ratio was more than
5% different from the terrestrial value (since almost all Solar System materials are within 5% of
terrestrial) at a significance of at least 2σ or if its N isotopic ratio was more than 3σ away from the
average value of DOM 08006 IOM (δ15N =10.5 ‰; Alexander et al., 2018). Seventeen 13C-rich
ROIs were identified as presolar SiC grains (e.g., Figure 3), based on the spatially correlated
presence of Si. An additional four highly 13C-rich grains were identified during the H- and C-
imaging runs (Section 3.1.3) and are probably SiC, but these measurements did not include Si and
the poorer spatial resolution for these measurements makes their true sizes and isotopic
compositions uncertain. The 17 SiC grains range in diameter from about 100 nm to 400 nm and
their measured 12C/13C ratios range from 27 to 79. These 12C/13C ratios are compared to those of
~16,000 SiC grains from the literature in Figure 8. The literature data are divided into those
measured as isolated single grains in meteorite acid residues (upper panel) and those measured in
situ as in this study (lower panel). Although the statistics are limited, the peak of the DOM 08006
SiC histogram plots at 12C/13C~70, compared to the peak of ~60 observed for the literature data
for isolated SiC grains. A similar shift is seen for the literature in situ data, indicating that it is
likely due to moderate isotopic dilution of the measured C isotopes by isotopically normal material
(e.g., contamination, epoxy used to prepare the meteorite section, or intrinsic organic matter).
The rest of the C- and/or N-anomalous regions are not spatially associated with Si and are most
likely particles of insoluble organic matter (IOM), though we cannot rule out that some are presolar
graphite grains (Haenecour et al., 2016). The identified C- and/or N-anomalous grains show a
similar range of grain sizes as the SiC grains and both enrichments and depletions are seen in 13C
and 15N (Table 3). The δ13C values of the C-anomalous grains of this study are compared to those
seen in the highly primitive CR chondrites QUE 99177 and MET 00426 (Floss and Stadermann,
2009b) and CO3.0 ALH 77307 (Bose et al., 2012) in Figure 9. The data for all of the meteorites
are consistent with two peaks in δ13C, with values of about ±150 ‰. However, we note that
typically grains with 13C/12C ratios within ~10% of terrestrial are by definition not counted as
11
anomalous in surveys such as ours. Thus, it is likely that the peaks in the histogram indicate that
the meteoritic organic grains mostly have δ13C values between -200 and +200 ‰, but only the
edges of the distribution are identified as anomalous.
The δ13C and δ15N values for the 36 C- and/or N-anomalous organic grains of this study for
which both elements were measured are shown in Figure 10. Also shown are the "normal" grains
from this study (those within 10% of terrestrial values within 3σ for both isotopic ratios) along
with data for similar anomalous grains measured in other primitive planetary materials. The C and
N data are not well correlated, and the DOM 08006 grains span similar ranges to those previously
seen in MET 00426 and ALH 77307. One difference is a lack of grains in DOM 08006 with δ15N
values higher than 1000 ‰, in particular grains with δ15N≈1000–1500 ‰ and depleted 13C (δ13C~
-200 to -100 ‰), as seen in other meteorites as well as in some IDP and comet Wild 2 samples
(dashed ellipse in Fig. 10).
One particularly interesting grain, A7_60, is highly depleted in both 13C and 15N (Table 3,
Figures 10 and 11). Grain A7_60 appears as two adjacent C-rich regions in NanoSIMS and SEM
images (Fig. 11), but the identical compositions of the two regions indicate that they were likely
originally part of the same inclusion, that was either split during the polishing of the thin section
or by the NanoSIMS beam, or is connected below the surface of the polished thin section. An
attempted measurement of the H isotopes in this grain was unsuccessful due to very low secondary
ion counts.
3.1.3. H-isotopic mapping
A total of 2,400 µm2 of Area 2 and 3,800 µm2 of Area 7 were mapped for H and C isotopes as
described in Section 2. Hydrogen was clearly correlated with C in all of the images as expected
both for IOM and the epoxy used to make the thin section. However, on average the D/H ratio of
the meteoritic C grains was indistinguishable from both that of the epoxy and of the terrestrial
standard we used and we found no evidence for spatial heterogeneity in D/H (Figure 12). DOM
08006 has a bulk δD value of about 0 ‰, but a purified IOM residue from this meteorite shows a
moderate bulk D enrichment (δD=476±25 ‰) and H/C=0.476 (Alexander et al., 2018). For our
imaging measurements, we obtained typical errors of ~200–500 ‰ for individual ~500-nm sized
carbonaceous grains and therefore we would not necessarily expect to resolve a 500 ‰ D
12
enrichment for an individual grain. However, it is somewhat surprising not to see even a bulk
enrichment when a large number of grains are averaged to obtain smaller statistical uncertainties.
Moreover, our results are in strong contrast to D/H imaging results from other primitive
carbonaceous chondrites, which have revealed highly heterogeneous D/H ratios, with δD values
in individual sub-µm grains (hotspots) reaching into the tens of thousands (Busemann et al., 2006;
Remusat et al., 2010). For example, in Figure 12 we compare NanoSIMS C and D/H images for
same-sized areas of DOM 08006 (panels a and c) and the primitive CR chondrite QUE 99177
(panels b and d). The measurement conditions were nearly identical for these images, as can be
seen from the similar 12C ion count rates (Figure 12a,c); H count rates (not shown) were likewise
similar. However, whereas QUE 99177 shows D hotspots with δD values ranging from 5,000 to
15,000 ‰, the total range of D/H in the DOM 08006 image is small and none of the apparent
isotopic variations are statistically significant. QUE 99177 IOM has a H/C=0.803 (Alexander et
al., 2007), much higher than that measured for DOM 08006 IOM. Since the two samples showed
similar H/C secondary ion ratios, we therefore conclude that the isotopically normal D/H we
measure on average in DOM 08006, compared to the known D-enriched composition of purified
IOM, is most likely due to H contamination in the sample. However, such contamination would
not be expected to mask hotspots with much higher intrinsic D/H ratios.
To investigate the apparent lack of D hotspots in DOM 08006, we obtained D/H images of
particles of IOM from both DOM 08006 and QUE 99177, pressed into gold foils as done in
previous work (e.g., Busemann et al., 2006). The imaging results for DOM 08006 IOM (Figure
12e) reproduce the bulk measured δD of ~500 ‰ and, as in the in situ results, show no evidence
for highly D-enriched hotspots, whereas these are abundant in the QUE 99177 data (Figure 12f).
The combined in situ and IOM results provide further support that the in situ H-isotope data are
affected by contamination and show that organic matter in DOM 08006 is much more isotopically
uniform in D/H than is seen in many other primitive meteorites.
3.2 Transmission Electron Microscopy
We successfully extracted four FIB sections containing five presolar silicate grains and
analyzed them by TEM. Note that the names of the grains have been changed since initial reports
13
of the TEM data (Stroud et al., 2013; Stroud et al., 2014), as indicated below. Compositional results
for the five grains are given in Table 4.
DOM-3 (formerly A2-18): Scanning TEM (STEM)-based EDS mapping revealed this Group
1 presolar grain to have an Fe-rich (Mg/(Mg+Fe)=0.38), non-stoichiometric composition with an
(Fe+Mg+Ca)/Si ratio of 1.5, which is intermediate between those of olivine and pyroxene (Table
4). No lattice fringes were identified in high-resolution TEM (HRTEM) imaging, which indicates
an amorphous structure, consistent with the intermediate composition.
DOM-8 (formerly A2C-15): A STEM-based EDS measurement indicates that this ~500-nm
Group 1 grain has a composition that is very close to stoichiometric olivine with a Mg/(Mg+Fe)
ratio of 0.83 (Table 4). We were unfortunately unable to make any structural measurements due to
the small volume of sample remaining after the SIMS measurement, and the overlap with adjacent
grains and protective C strap. However all previous presolar silicates with olivine compositions
for which structural information is available were found to be crystalline olivine (e.g., Messenger
et al., 2005; Busemann et al., 2009; Vollmer et al., 2009a). We thus consider it most likely that
this grain is indeed forsteritic olivine with Fe contents within the range previously seen for presolar
grains.
DOM-17,18 (formerly A2C2-25a,b): DOM-17 and DOM-18 were located 4 µm apart in the
meteorite section and we were thus able to prepare a single FIB section including both grains
(Figure 13). DOM-17 is the most 17O-rich gain we identified in DOM 08006, though its
composition still lies within the Group 1 field (Fig. 6). In contrast, the neighboring DOM-18 grain
is the most 18O-poor grain we identified (Fig. 6) and although its composition lies within the Group
1 field, it may well be a Group 2 grain for which a small amount of normal O from surrounding
material diluted its anomalous 18O/16O ratio. STEM-based EDS mapping (Table 4) indicates an
Fe-rich, non-stoichiometric composition for DOM-17, similar to that of DOM-3, and a Mg-rich
olivine composition for DOM-18 (Fig. 14). DOM-17 is amorphous, whereas a selected-area
diffraction pattern of grain DOM-18 indexes to crystalline olivine (Fig. 14), in agreement with its
chemical composition.
DOM-77 (formerly A3C-CN1): STEM-based EDS mapping (Figure 15) shows that grain
DOM-77 is compositionally zoned, with a Mg-silicate rim surrounding inner zones of Ca- and Al-
14
rich material. HRTEM imaging reveals that the Mg-silicate region is polycrystalline with lattice
fringe spacings consistent with forsteritic olivine. Lattice fringes from the Ca-rich zone index to
hibonite, anorthite, or monticellite, but not to grossite, gehlenite, or akermanite. No lattice fringes
definitively associated with the Al-rich zone were identified. The olivine at the top surface of the
grain, where it is most easily distinguished from adjacent matrix, has a Mg/(Mg+Fe)≈0.93. The
high Ca/Al ratio (0.5, Table 4) indicates that hibonite is not the dominant Ca-bearing phase.
However, the compositions of the distinct Ca- and Al-rich sub-grains could not be accurately
determined because they are smaller than the thickness of the FIB section. Overall, the mineralogy
of this composite grain is reminiscent of the larger amoeboid olivine aggregates (AOAs) present
in chondrites (Scott and Krot, 2003) and IDPs (Joswiak et al., 2013).
4.1. Presolar Grains
4.1.1. Isotopic compositions
4. Discussion
The O-isotopic ratios observed in the 101 presolar O-rich grains (Figures 6 and 7) span similar
ranges to those observed before for presolar oxides and silicates, and fall within the four Group
definitions of Nittler et al. (1997). As discussed in the literature, the 17O-rich, 18O-poor
compositions of the largest population of grains (Group 1) point to an origin in asymptotic giant
branch (AGB) stars (Huss et al., 1994; Nittler et al., 1994; Nittler et al., 1997; Nittler et al., 2008;
Nittler, 2009). The highly 18O-depleted Group 2 grains have long been considered to have
originated in low-mass (<2 M) AGB stars undergoing a poorly-constrained extra mixing process
(Wasserburg et al., 1995; Nollett et al., 2003; Palmerini et al., 2017), but a recent measurement of
a key nuclear reaction rate suggests that they may instead have formed around more massive,
intermediate-mass AGB stars, 4–8 M (Lugaro et al., 2017). The 18O-enriched Group 4 grains
most likely formed in Type II supernovae (Choi et al., 1998; Nittler et al., 2008), whereas the
Group 3 grains, with sub-solar 17O/16O and normal to depleted 18O/16O ratios, likely formed in
either supernovae or low-metallicity AGB stars. A small fraction of presolar grains, with
17O/16O>0.004, may have originated in classical novae (Gyngard et al., 2010; Leitner et al., 2012a),
but none of the DOM 08006 grains show such compositions. The implications of the grain isotopic
15
signatures for stellar and galactic evolution have been discussed extensively elsewhere (Nittler et
al., 2008; Nittler, 2009) and we do not repeat those discussions here.
The 12C/13C ratios of the DOM 08006 presolar SiC grains (Table 3 and Figure 8) lie in the
range of the dominant "mainstream" population, which makes up ~90% of all presolar SiC (Zinner,
2014) and are thought to have formed around C-rich AGB stars (e.g., Hoppe et al., 1994; Lugaro
et al., 2003). Statistically, all of the DOM 08006 SiC grains are almost certainly mainstream grains,
although without additional isotopic data we cannot rule out some belong to the rare X or Z classes
(≤2% of SiC) as these grains can also have 12C/13C ratios in the observed range.
4.1.2. Abundances
The matrix-normalized abundances of O-rich presolar grains in the four analyzed areas are
given in Table 1. These were calculated by dividing the total area covered by presolar grains by
the total area analyzed in the imaging runs (excluding large cracks, etc). To be conservative, we
used the grain sizes corrected for the primary beam diameter (Table 2, Figure 4). We assume that
the average density of the grains is similar to that of the matrix in order to convert areal density to
mass concentration. Three of the four areas show consistently high abundances of 200–300 ppm,
whereas Area 7 shows a lower abundance of 155 ppm. However, as discussed earlier, the primary
beam was accidentally slightly defocused for this run, leading to poorer spatial resolution and
hence poorer detection efficiency for presolar O-rich grains. We thus calculate the average O-rich
presolar grain abundance in DOM 08006 to be 257 ppm by combining the data from Areas 2, 3,
and 6 (Table 1). Note that without our correction for the primary beam broadening, the inferred
abundance would increase to about 300 ppm.
Essentially all published studies have estimated uncertainties on presolar grain abundances
simply from Poisson statistics based on the number of identified grains (often using the tables of
Gehrels, 1986). However, this method ignores possible contributions to the error budget from the
wide range of grain sizes and the uncertainty in determining individual grain sizes. To investigate
this further, we used a Monte Carlo method to estimate the uncertainty in our abundance estimate.
We calculated 50,000 simulated grain distributions based on the 90 presolar grains identified in
Areas 2, 3, and 6. For each instance, we simulated N grains, where N was taken from a Poisson
16
distribution with a mean value of 90. Then, the size of each of the N grains was chosen by selecting
a random grain from the DOM 08006 data set, of size n pixels, and assigning it n*' pixels from a
Poisson distribution with average value n. Finally, the size of each simulated grain was corrected
for a 120-nm primary beam size as described above for the real data, and a total abundance
calculated. The resulting distribution of simulated abundances is shown in Figure 16. We estimate
1σ and 2σ uncertainties based on the limits enclosing 84.1 % and 97.7 % of the simulations,
respectively (Table 1 and Figure 16). For comparison, simply using the Gehrels (1986) tables for
90 grains would imply 1σ uncertainty intervals about our average value of 257 ppm of +30 and -
27, compared to the values of +44 and -41 that we estimate from the Monte Carlo simulations. It
is thus likely that previous reports of presolar silicate abundances in meteorites have
underestimated the statistical uncertainties by ignoring the effect of variable grain sizes.
In Figure 17, we compare our derived matrix-normalized abundance of presolar O-rich grains
in DOM 08006 to other primitive meteorites, Antarctic micrometeorites, and IDPs. Abundances
are shown with 1σ error bars (as reported by original publications) and plotted against the total
analyzed area of each sample. We presume that data acquired from larger areas are more
statistically robust. Note that we do not include the recent data of Hoppe et al. (2015, 2017). These
studies were made with higher NanoSIMS spatial resolution and hence better sensitivity, and thus
cannot be directly compared to the data discussed here, all of which was acquired under similar
analytical conditions. Our estimates of the presolar silicate and oxide grain abundance in DOM
08006 are higher than those reported in the other primitive carbonaceous chondrites and Antarctic
micrometeorites, but still lower than estimated for IDPs, although the 2σ lower limit is comparable
to the value of ~180 ppm seen for several primitive meteorites, and the 2σ upper limit overlaps
with the error limit on the average abundance in IDPs. Note that to construct Figure 17 we
combined two reported sets of abundances for both QUE 99177 (Floss and Stadermann, 2009a;
Nguyen et al., 2010) and MET 00426 (Floss and Stadermann, 2009a; Leitner et al., 2016) to
decrease the statistical errors. For both meteorites, the Floss and Stadermann (2009a) study found
abundances that were greater than 200 ppm, overlapping with our estimate for DOM 08006, but
in the case of QUE 99177, analysis of more than twice as much area by Nguyen et al. (2010)
indicated a significantly lower abundance for this meteorite. Moreover, the latter study was
performed with the same instrument, protocols and analysis software as the present work and is
17
thus directly comparable. The evidence, therefore, suggests that DOM 08006 has a higher
abundance of presolar O-rich grains in its matrix than does QUE 99177, though the estimates
overlap at the 2σ level.
Haenecour et al. (2018) have recently reported presolar grain surveys in DOM 08006, ALH
77307, and the CO3.0 chondrite LaPaz Icefield (LAP) 031117. They found lower abundances in
fine-grained chondrule rims compared to interchondrule matrix, but since our study and almost all
others have focused on matrix we only show their matrix abundances in Fig. 17. The DOM 08006
matrix abundance found by Haenecour et al. (2018), 227±32 ppm, agrees very well with ours. We
thus combined the two data sets to arrive at a weighted average abundance of O-rich presolar grains
in DOM 08006 of 240±30 ppm (Fig. 17). This 1σ error is estimated from counting statistics of the
total of 141 grains, and multiplied by 1.5, based on the Monte Carlo results described above.
The data shown in Figure 17 provide strong support that the matrix of DOM 08006 has an
abundance of presolar silicates and oxides that is higher than any other carbonaceous chondrite
studied to date and is thus a highly primitive meteorite. We note that a comparably high abundance
of presolar silicates (275±50 ppm, based on 32 grains) has been reported for an ordinary chondrite,
Meteorite Hills 00526, in a conference abstract (Floss and Haenecour, 2016a). When we compare
our results for DOM 08006 to those of QUE 99177 and ALH 77307, acquired in our laboratory by
means of very similar analytical techniques (Nguyen et al., 2010), it is apparent that presolar grains
are better preserved in DOM 08006. The ratio of presolar silicate grains to presolar oxide grains
has also been taken as a sensitive indicator of alteration in a given meteorite, as silicate minerals
are more susceptible to destruction by thermal metamorphism or aqueous alteration than are oxide
minerals (Floss and Stadermann, 2009a; Leitner et al., 2012b). As discussed above in Section
3.1.1., our data indicate a lower limit on this ratio of 4.4, but the true ratio is likely much higher
and indeed Haenecour et al. (2018) reported a ratio of 24.5 for this meteorite, though their
statistical uncertainty is high as their estimate is based on 49 silicate grains and 2 oxides. Leitner
et al. (2012b) made a simple estimate based on elemental abundances that the silicate/oxide dust
ratio in stardust forming around AGB stars (the dominant sources of the presolar O-rich grains)
should be ~23, similar to the average value of 22 they estimated for primitive CR chondrites and
IDPs. The high apparent ratio in DOM 08006 is further evidence that this is a highly primitive
meteorite.
18
We calculated the matrix-normalized abundance of SiC in DOM 08006 by the same
methodology discussed above for presolar O-rich grains, with the exception that we included Area
7 in the calculation. The slightly poorer spatial resolution of the Area 7 measurements has a smaller
effect on the detection efficiency of SiC than of O-rich phases since there is much less C
surrounding the grains to dilute the isotopic anomalies. We obtained an average abundance of 35
ppm. Monte Carlo modeling like that shown for O-rich grains in Figure 16 indicates 1σ errors of
±10 ppm about the average and 2σ errors of +25 and -17 ppm; that is, at a 97.7% certainty level,
the abundance lies between 18 ppm and 60 ppm. Davidson et al. (2014a) used NanoSIMS ion
imaging of IOM residues to estimate presolar SiC abundances in members of various primitive
meteorite groups. They found that CI chondrites, most CR chondrites, and the highly primitive
ungrouped Acfer 094 all have a roughly constant abundance of presolar SiC of ~30 ppm. Leitner
et al. (2012b) found a higher abundance of 131±53 ppm in the CR2 NWA 852, but this value is
within uncertainty of the Davidson et al. (2014a) CR data. In contrast, both in situ NanoSIMS and
noble gas data imply a lower abundance of SiC of ~10 ppm in the CO3.0 ALH 77307 (Huss et al.,
2003; Davidson et al., 2014a). Davidson et al. (2014a) interpreted the constant SiC abundance of
~30 ppm seen for many primitive meteorites as indicating a uniform abundance in the chondrite-
forming region of the solar nebula, with the lower abundances seen in some meteorites (e.g., ALH
77307) reflecting partial grain destruction by parent body processing such as thermal
metamorphism. More recently, by combining their data with those of Bose et al. (2012), Haenecour
et al. (2018) found a higher abundance of 59±13 ppm for C-anomalous grains (assumed to be
mostly SiC) in ALH 77307, consistent within errors with the Davidson et al. (2014a) value of 30
ppm. The source of the discrepancy between this value and the lower ones reported for ALH 77307
by Huss et al. (2003) and Davidson et al. (2014a) is unclear, but may indicate that this meteorite
experienced heterogeneous thermal metamorphism on relatively small scales. All in all, our SiC
abundance for DOM 08006 is clearly in very good agreement with the average seen in most
primitive chondrites, supporting the idea that different chondrite classes accreted matrix with a
constant amount of presolar SiC. We note that we found an uneven distribution of SiC grains in
the four analyzed areas of the thin section (Table 1), with five to seven identified grains in each of
Areas 2, 3, and 7, and zero in Area 6. However, the total imaged area of Area 6 was considerably
smaller than the other areas and the non-detection of any SiC grains is statistically consistent with
a 35 ppm average abundance.
19
4.1.3. Mineralogy of presolar silicates
Most astronomical knowledge of circumstellar dust mineralogy is based on infrared
spectroscopy, which has revealed a range of amorphous and crystalline silicate and oxide phases
around O-rich AGB stars (e.g., Molster et al., 2010). However, stellar infrared spectra reflect
averages over large numbers of individual grains and are often difficult to unambiguously
interpret. In this regard, presolar grains are a complementary and valuable tool for understanding
the range of compositions and structures of AGB stardust and thus for better understanding stellar
dust formation processes.
As discussed in Section 3.2 we obtained TEM data on fives presolar silicate grains (Figs. 13-
15, Table 4). Two grains are amorphous and non-stoichiometric with Fe-rich compositions
intermediate between those of pyroxene and olivine, two grains are Mg-rich olivine (tentative for
DOM-8), and one grain-DOM-77-is a composite with Al and Ca-rich material surrounded by
nanocrystalline forsteritic olivine. TEM data have now been reported for more than 50 presolar
silicates from meteorites and IDPs (reviewed in Floss and Haenecour, 2016b). Also, reliable
major-element chemical data from Auger spectroscopy have been reported for more than 400
presolar silicate grains (e.g., Floss and Stadermann, 2009a; Stadermann et al., 2009; Nguyen et al.,
2010; Bose et al., 2012; Floss and Haenecour, 2016b; Haenecour et al., 2018). The compositions
and structures of the four non-composite grains analyzed here are within the range of previous
observations of presolar silicates in general and in CO3.0 chondrites specifically. For example,
Haenecour et al. (2018) classified presolar silicates on the basis of their (Fe+Mg+Ca)/Si ratios and
found that grains with compositions that are olivine-like or intermediate between olivine and
pyroxene, like the grains reported here, each make up roughly a quarter of the presolar grains in
CO3.0 chondrites. Moreover, the two non-stoichiometric amorphous grains reported here, DOM-
3 and DOM-17, are Fe-rich, with atomic Fe/Mg ratios of 1.6 and 2.6 and Fe contents of 14 at.%
and 17 at.%, respectively. This Fe-richness is typical of presolar silicates; two-thirds of the grains
from CO3.0 chondrites analyzed by Auger have Fe/Si>1.2 and the median Fe contents for CO3.0
presolar grains is 17 at.% (Haenecour et al., 2018). We did not identify any pyroxene-like or SiO2-
rich or –poor grains, like those seen in the Auger studies, but our TEM statistics are far too limited
to make realistic comparisons with the much larger Auger datasets. Finally, considering the
20
available TEM data for presolar silicates as a whole, roughly two-thirds of the analyzed grains
from both chondrites and IDPs are amorphous and the rest are at least partially crystalline. Again,
our statistics are too small to make quantitative conclusions, but the observation of both amorphous
and crystalline grains in our limited data set appears to be consistent with the crystalline fraction
seen in the larger data set. All in all, the four non-composite presolar silicates from DOM 08006
analyzed by TEM are not atypical of the population of presolar silicates as a whole.
In contrast, composite grains like DOM-77 are rare among the presolar silicate population, but
the compositions and structures of such grains likely capture changing conditions (e.g.,
temperature, composition) in the stellar gases from which they condensed and thus in principle can
provide clues to grain formation in stellar environments. A few other presolar grains containing
Ca-Al-rich phases co-existing with ferromagnesian silicates have been reported, including an
aggregate of 30-nm scale Ca or Al oxides interspersed with Mg silicates (Nguyen et al., 2010), an
amorphous Ca-Si-rich grain with embedded nanocrystals of hibonite (Vollmer et al., 2013), a
MgAl2O4 spinel surrounded by a nonstoichiometric amorphous Mg-rich silicate (Nguyen et al.,
2014), an oxide-silicate aggregate with hibonite and Mg-rich olivine composition (Floss and
Stadermann, 2012), and a large Al-Ca-Ti oxide grain mantled by a Ca-rich silicate with a
pyroxene-like composition (Leitner et al., 2017). Note that no TEM data are available for the latter
two grains, so there are no constraints on the microstructures of the various phases included. Of
these, grain DOM-77 is somewhat reminiscent of the hibonite-olivine composite reported by Floss
and Stadermann (2012), though that object did not have the core-mantle structure of DOM-77.
Moreover, as discussed above in Section 3.2, the core of DOM-77 is too Ca-rich to be hibonite,
though the TEM measurements did not fully resolve the substructure of the core.
The composite structure of a grain with a core enriched in refractory elements like Ca and Al,
mantled by less refractory Mg-rich silicate, is highly suggestive of condensation from a cooling
gas. That is, in a high-temperature (>1300 K) gas of solar composition (a good approximation for
O-rich AGB star envelopes) under thermodynamic equilibrium, oxides and silicates rich in Al, Ca,
and Ti (e.g., corundum, hibonite, perovskite, grossite, melilite, etc.) condense, followed by Mg-
rich silicates (forsterite followed by enstatite), iron metal, and other phases at lower T as the earlier
formed phases react with the ambient gas. However, models of circumstellar dust formation in
AGB outflows suggest that the process is not fully controlled by equilibrium condensation, with
21
the kinetics of grain nucleation and growth playing a key role (Höfner et al., 2009; Gail, 2010). As
direct samples of AGB dust production, presolar grains obviously can provide important clues to
the process. TEM studies of refractory Al-rich presolar oxide grains (e.g., corundum, spinel, and
hibonite) have revealed most, but not all, to be well-crystallized close-to-stoichiometric minerals
that probably formed by equilibrium condensation at high temperature (Stroud et al., 2004; Zega
et al., 2011; Takigawa et al., 2014; Zega et al., 2014) and a similar conclusion can be drawn for
presolar SiC (Daulton et al., 2003). This suggests that grain condensation at the highest
temperatures in AGB stellar envelopes is governed by thermodynamic equilibrium, even though
kinetics must play a critical role in the nucleation and growth of circumstellar dust grains
(Bernatowicz et al., 1996; Gail and Sedlmayr, 1999). In contrast, the dizzying range of non-
stoichiometric compositions and microstructures observed for presolar silicates (Floss and
Haenecour, 2016b) and Fe-Cr-rich spinels (Zega et al., 2014) indicates formation far from
equilibrium, likely dominated by kinetic effects.
Core-mantle grains like DOM-77 and the others described above provide the opportunity to
probe the transition from equilibrium to non-equilibrium condensation in stellar outflows. For
example, the mantling of a crystalline spinel grain by amorphous silicate seen by Nguyen et al.
(2014) strongly suggests such a transition in the parent star of that presolar grain. Yet the
crystalline nature of both the Al-Ca-rich core phases and the Mg-rich olivine rim of DOM-77
suggests that, in contrast, equilibrium conditions prevailed through the full history of the grain's
condensation. In fact, the composition of this grain is similar to that of amoeboid olivine aggregates
in primitive meteoritic materials (Scott and Krot, 2003; Joswiak et al., 2013), widely considered
to be equilibrium condensates from the solar nebula (Grossman and Steele, 1976). A similar
argument may be made for the oxide-silicate grain reported by Floss and Stadermann (2012), but
the crystallinity of that grain is unknown. A detailed study of grain formation in stellar
environments to quantify what conditions may have led to the observed properties of DOM-77 is
beyond the scope of this paper, but this result clearly adds to the evidence of highly variable and
dynamic conditions during the formation of dust around AGB stars.
22
4.2. Organic matter
The C NanoSIMS images acquired here revealed abundant sub-µm carbonaceous inclusions
dispersed throughout the matrix of DOM 08006 (Figure 3; Section 3.1.2), similar to what has been
seen in SIMS studies of other carbonaceous chondrites (Nakamura-Messenger et al., 2006;
Remusat et al., 2010; Bose et al., 2012; Peeters et al., 2012; Floss et al., 2014; Le Guillou et al.,
2014; Alexander et al., 2017). Although it is difficult to ascertain the original morphology of the
grains from SEM images, at least one of the 15N-enriched grains appears to be a nanoglobule (Fig.
3; Garvie and Buseck, 2004; Nakamura-Messenger et al., 2006; Floss and Stadermann, 2009b),
providing further evidence that these are a common constituent of organic matter in all classes of
primitive extraterrestrial materials including chondrites, micrometeorites, IDPs and comet Wild-2
samples (Dobrică et al., 2009; De Gregorio et al., 2010; Matrajt et al., 2012; De Gregorio et al.,
2013). De Gregorio et al. (2013) found evidence for two chemical populations of nanoglobules in
meteoritic IOM, one with similar C functional groups to the non-globular IOM and one showing
a distinctly more aromatic character. Unfortunately, we do not have functional-group data for
nanoglobules in DOM 08006 to make a direct comparison.
The origin of the IOM in extraterrestrial materials is unclear and controversial (Alexander et
al., 2017), although the most extreme H and N isotopic anomalies are commonly taken as evidence
for an interstellar provenance for at least a portion of the matter (Messenger, 2000; Busemann et
al., 2006). This reflects astronomical observations of interstellar molecules showing a wide range
of D/H and 15N/14N ratios as well as the results of theoretical astrochemical models predicting
large fractionations in molecular clouds due to a variety of processes, including for example low-
temperature gas-phase ion molecule and gas-grain chemical reactions (Ehrenfreund and Charnley,
2000; Sandford et al., 2001). Carbon isotopic anomalies are much rarer in meteoritic organics, but
they are also predicted by astrochemical modeling and interstellar chemistry has thus been invoked
to explain these as well (Floss and Stadermann, 2009b; Bose et al., 2014). The observation of
similar C- and N-isotopically anomalous sub-µm organic grains in the CO3s DOM 08006 and
ALH 77307 (this work; Bose et al., 2014), CRs (Busemann et al., 2006; Floss and Stadermann,
2009b), IDPs of possible cometary origin (Floss et al., 2004) and comet Wild 2 samples
23
(McKeegan et al., 2006; De Gregorio et al., 2010) suggests that such grains were distributed
throughout the protosolar disk.
Meteoritic organic matter is chemically and isotopically heterogeneous on all scales, from
individual sub-µm grains as studied here, to bulk meteorites. A key unanswered question regarding
meteoritic organics is whether differences in bulk organic properties between chondrite groups
reflects divergent parent-body processing of a common precursor material or that the organics
accreted by different groups was fundamentally different. Data from different portions of the
primitive ungrouped carbonaceous chondrite Tagish Lake show that much of the variability in the
chemical and isotopic properties of both soluble and insoluble organics in CI, CM and CR
chondrites is probably due to aqueous alteration in the meteorites' parent asteroids (Herd et al.,
2011; Alexander et al., 2014). The IOM in CR chondrites and the unusual CM chondrite Bells
bears many similarities to that found in IDPs and inferred for comet Halley and is probably the
most primitive of any meteoritic IOM (Alexander et al., 2007). In contrast, with the possible
exception of LL3.00 Semarkona, the IOM in even the most primitive members of chondrite groups
that have experienced thermal metamorphism (e.g., unequilibrated ordinary, CO3, and CV3
chondrites) appears to have experienced some heating relative to that found in CRs, based on
compositional data, Raman spectroscopy, and X-ray absorption spectroscopy (Bonal et al., 2006;
Alexander et al., 2007; Busemann et al., 2007; Cody et al., 2008; Bonal et al., 2016) and its
properties can be a sensitive probe of the degree of metamorphism.
Bonal et al. (2016) found that Raman spectral parameters of organic matter in DOM 08006
indicate less heating than that experienced by ALH 77307, previously thought to be the most
primitive CO3.0 chondrite. This conclusion is supported by the data of Alexander et al. (2018),
who found that DOM 08006 has a higher bulk C abundance, and its IOM has higher H/C, D/H and
15N/14N ratios than ALH 77307. Because the H/C ratio of IOM in DOM 08006 is similar to that in
the unequilibrated ordinary chondrite (LL3.00) Semarkona, Alexander et al. (2018) suggested that
the two meteorites experienced similar thermal histories and that DOM 08006 should thus be
classified as a 3.00. As discussed above in Section 3.1.3, our NanoSIMS measurements revealed
no resolvable heterogeneity in D/H ratios in DOM 08006, either in situ or in isolated IOM,
consistent with loss due to metamorphism (Alexander et al., 2010). The lower D/H and 15N/14N
values for bulk IOM from DOM 08006 compared to primitive CR chondrites indicate that the
24
metamorphism has affected both elements, but the preservation of 15N-rich hotspots as well as C-
anomalous grains in DOM 08006 indicate that the molecular carriers of these anomalies are more
thermally resistant than those of D anomalies within primitive IOM. Thus, it appears that D-rich
moieties in IOM are less resistant both to thermal metamorphism and to aqueous alteration (Herd
et al., 2011) on parent bodies than are carriers of N anomalies. Moreover, the loss of D-rich, and
to a lesser extent, 15N-rich, organic molecules without significantly affecting other inorganic signs
of primitiveness (e.g., the high presolar grain abundances reported here; the high Cr contents of
olivine reported by Davidson et al., 2014b) provides further evidence that IOM properties are
among the most sensitive probes of metamorphism in highly primitive chondrites. In this regard,
the increasing recognition that organics can be used as sensitive probes of parent-body processing
suggests that a new petrologic scale, including consideration of organic properties, may be needed
to better describe the nature of highly primitive chondrites. However, this would require a better
understanding than currently exists of the response of organic matter to parent-body processing
under a wide range of conditions (see, e.g., the discussion of difficulties arising when attempting
to define a Raman-based C thermometer for Type 3 chondrites in Bonal et al., 2016).
Based on the total areas of identified anomalous grains relative to the total mapped areas, and
correcting for an assumed factor of two lower density of IOM compared to the matrix as a whole,
we estimate abundances of ~33 ppm and ~135 ppm for C-anomalous and N-anomalous grains,
respectively; grains anomalous in both elements are counted in both abundance estimates.
However, if we limit consideration to the combined C- and N-isotopic imaging runs of matrix Area
3 ("A3cCN" grains in Table 3), we find a much higher abundance of C-anomalous grains (~400
ppm). This higher abundance may indicate true heterogeneity in the meteorite, but more likely
reflects differences in image analysis methodology (automatic definition of C-rich regions of
interest as opposed to ratio image generation). Thus, the data are consistent with spatial
heterogeneity in the abundance of C-anomalous grains, but additional systematic C and N mapping
would be required to tell this definitively. In comparison, Floss and Stadermann (2009b) found an
abundance of C-anomalous grains in the CR chondrites QUE 99177 and MET 00426 of ~120 ppm
and Bose et al. (2012) reported an abundance of 15N-rich grains in CO3.0 ALH 77307 of 160±30
ppm. Neither study provided details of how the abundances were estimated (e.g., whether or not a
density correction was applied to convert area fraction to mass fraction). Nevertheless, it is
apparent that the abundances of C-anomalous and 15N-rich grains in DOM 08006 are roughly
25
similar to those in the least altered CR chondrites, providing further support that this is a highly
primitive meteorite.
Note that Bose et al. (2014) reported a remarkably high abundance (~900 ppm) of very 15N-
rich grains, with δ15N values ranging from 1600 to 3000, in the CO3 chondrite QUE 97416. These
authors further identified a single presolar silicate grain, with an inferred presolar silicate
abundance of 6 ppm, indicating that this meteorite is likely more processed than either DOM 08006
or ALH 77307. Given this, the high abundance of extremely 15N-rich grains is puzzling; similar
grains were not identified here and are absent or very rare in other chondrites (e.g., Figure 10).
Possible explanations are that QUE 97416 originated on a different parent body than other CO3s,
one that sampled a different population of anomalous organic grains, that thermal metamorphism
or some other process somehow enriched some organic grains in 15N in QUE 97416, or that the
data reported by Bose et al. (2014) are compromised by an unrecognized instrumental artifact.
As mentioned above in Section 3.1.2, the NanoSIMS mapping revealed a grain, A7_60
(Figures 10 and 11), with large depletions in both 13C and 15N relative to terrestrial isotopic
compositions (δ13C=-300 ‰, 12C/13C=127, δ15N=-264 ‰, 14N/15N=371). These isotopic
compositions are similar to those of some rare presolar SiC and graphite grains thought to come
from low-metallicity AGB stars (e.g., Figs. 3 and 12 of Zinner, 2014); SEM-EDS analysis only
revealed C for this grain raising the possibility that it is indeed a presolar graphite that got sheared
into two fragments during polishing of the DOM 08006 thin section. However, the NanoSIMS
measurement of this grain revealed it to be rich in N, with a measured CN-/C- secondary ion ratio
of ~0.8, roughly corresponding to ~10% N (Alleon et al., 2015), suggesting that it is in fact N-rich
organic matter and not graphite. This does not rule out it being a presolar circumstellar organic
grain, as organic molecules are indeed observed around evolved stars (Matsuura et al., 2014), but
it does have a relatively unusual isotopic composition for AGB stardust. Alternatively, the N-
isotopic composition of grain A7_60 is close to, but not as 14N-rich as, that of the solar wind (δ15N
≈ -400 ‰; Marty et al., 2011), which is assumed to be representative of the bulk protosolar
composition. The C-isotopic composition of the Sun is unknown. Hashizume et al. (2004) argued
on the basis of SIMS depth profile measurements of lunar soils that it is isotopically light relative
to the Earth, with δ13C <-100 ‰. Recently, Lyons et al. (2017) reported a spectroscopic
measurement of CO in the solar photosphere, yielding δ13C = -48 ± 7‰. Thus, although the lunar
26
measurement of Hashizume et al. (2004) does not rule out as extreme a 13C depletion in the Sun
as observed in A7_60, the astronomical measurement argues against it. We suggest that the most
likely origin of grain A7_60 is in the Sun's parental molecular cloud, where its C isotopic
composition was set by interstellar fractionation processes that were decoupled from those
affecting N isotopes (e.g., Langer et al., 1984), with the latter mainly reflecting the bulk isotopic
composition of the cloud.
5. Conclusions
A detailed NanoSIMS and TEM study of a polished thin section of the CO3.0 chondrite DOM
08006 allows us to draw the following conclusions:
1) NanoSIMS isotopic mapping revealed that DOM 08006 has the highest abundance of
presolar O-rich (silicate and oxide) grains in its matrix, 257 +76/-96 ppm (2σ), found thus
far in any carbonaceous chondrite, in good agreement with an independent measurement
of a different thin section by Haenecour et al. (2018). Combining our dataset with that of
Haenecour et al. (2018) gives an average abundance of 240±30 ppm (1σ). The NanoSIMS
data also revealed a SiC abundance of 35 (+25/-17, 2σ) ppm, similar to that seen in other
highly primitive meteorites. Within the statistical uncertainty of our observations, the
abundance of presolar grains in DOM 08006 is homogeneous on a scale of several mm.
The distributions of isotopic compositions of the presolar grains identified here are
consistent with previous observations in other primitive meteorites with differences being
fully explainable by analytical effects (e.g., dilution of isotopic signatures by surrounding
material during NanoSIMS measurements).
2) Transmission electron microscopy was performed on five presolar silicates. One grain was
found to have a composite mineralogy that is similar to larger amoeboid olivine aggregates
found in chondrites, and consistent with condensation near thermodynamic equilibrium. In
addition, the TEM observations revealed two non-stoichiometric amorphous grains and
two olivine grains, though one of the latter is identified as such solely based on its
composition.
3) NanoSIMS C and N mapping revealed IOM to be present primarily as sub-micron
inclusions with ranges of C- and N-isotopic anomalies similar to those seen in primitive
27
CR chondrites and interplanetary dust particles, but with normal (terrestrial) and
homogeneous D/H ratios. Most likely, DOM 08006 and other CO chondrites accreted a
similar complement of primitive and isotopically anomalous organic matter to that found
in other chondrite classes and IDPs, but the very limited amount of thermal metamorphism
experienced by DOM 08006 has caused loss of D-rich organic moieties, while not
substantially affecting the molecular carriers of C and N anomalies or the inorganic phases
of the meteorite. In this regard, the D/H of organic matter may be the most sensitive probe
available of thermal metamorphism in highly primitive chondrites. One organic grain that
was highly depleted in 13C and 15N was identified; we propose it originated in the Sun's
parental molecular cloud.
4) Together with other isotopic and petrographic evidence (Davidson et al., 2014b; Bonal et
al., 2016; Alexander et al., 2018), our data indicate that DOM 08006 has experienced the
least parent body modification of any known CO chondrite. Compared to other highly
primitive carbonaceous chondrites (e.g., CRs QUE 99177 and MET 00426, and the
ungrouped Acfer 094), it contains a higher abundance of O-rich presolar grains in its matrix
as well as organic matter with similar ranges of C and N isotopic anomalies, but more
homogeneous D/H ratios. These differences most likely reflect the effects of the limited
parent body processing under variable conditions that these meteorites experienced (e.g.,
aqueous alteration for CRs and Acfer 094, heating for DOM 08006). The fact that no
uniquely primitive single meteorite has yet been identified with no signs of parent body
processing highlights the need for continued comparative work on all of these samples to
help further elucidate the earliest stages of parent body processing on volatile-rich
asteroids.
Acknowledgements: The authors would like to thank Cecilia Satterwhite and Kevin Righter
(NASA, Johnson Space Center) for supplying the section of DOM 08006. US Antarctic meteorite
samples are recovered by the Antarctic Search for Meteorites (ANSMET) program, which has
been funded by NSF and NASA, and characterized and curated by the Department of Mineral
Sciences of the Smithsonian Institution and the Astromaterials Curation Office at NASA Johnson
Space Center. We thank the Associate Editor, Eric Quirico, and the referees, Christine Floss and
28
Henner Busemann for constructive and helpful comments. We also acknowledge Ryan Ogliore for
inspiring the Monte Carlo error estimation method used in this paper. This work was supported in
part by NASA grants NNX10AI63G and NNX11AB40G.
References
Alexander, C.M.O'D., Cody, G.D., Gregorio, B.T.D., Nittler, L.R. and Stroud, R.M. (2017) The
nature, origin and modification of insoluble organic matter in chondrites, the major source of
Earth's C and N. Chemie der Erde - Geochemistry 77, 227-256.
Alexander, C.M.O'D., Cody, G.D., Kebukawa, Y., Bowden, R., Fogel, M.L., Kilcoyne, A.L.D.,
Nittler, L.R. and Herd, C.D.K. (2014) Elemental, isotopic, and structural changes in Tagish
Lake insoluble organic matter produced by parent body processes. Meteorit. Planet. Sci. 49,
503–525.
Alexander, C.M.O'D., Fogel, M., Yabuta, H. and Cody, G.D. (2007) The origin and evolution of
chondrites recorded in the elemental and isotopic compositions of their macromolecular
organic matter. Geochim. Cosmochim. Acta 71, 4380–4403.
Alexander, C.M.O'D., Greenwood, R.C., Bowden, R., Gibson, J.M., Howard, K.T. and Franchi,
I.A. (2018) A multi-technique search for the most primitive CO chondrites. Geochim.
Cosmochim. Acta 221, 406-420.
Alexander, C.M.O'D., Howard, K.T., Bowden, R. and Fogel, M.L. (2013) The classification of
CM and CR chondrites using bulk H, C and N abundances and isotopic compositions.
Geochim. Cosmochim. Acta 123, 244–260.
Alexander, C.M.O'D., Newsome, S.D., Fogel, M.L., Nittler, L.R., Busemann, H. and Cody, G.D.
(2010) Deuterium enrichments in chondritic macromolecular material--Implications for the
origin and evolution of organics, water and asteroids. Geochim. Cosmochim. Acta 74, 4417–
4437.
Alexander, C.M.O'D., Nittler, L.R., Davidson, J. and Ciesla, F.J. (2017c) Measuring the level of
interstellar inheritance in the solar protoplanetary disk. Meteorit. Planet. Sci. 52, 1797-1821.
Alleon, J., Bernard, S., Remusat, L. and Robert, F. (2015) Estimation of nitrogen-to-carbon ratios
of organics and carbon materials at the submicrometer scale. Carbon 84, 290-298.
Bernatowicz, T.J., Cowsik, R., Gibbons, P.C., Lodders, K., Fegley, B., Jr., Amari, S. and Lewis,
R.S. (1996) Constraints on stellar grain formation from presolar graphite in the Murchison
meteorite. Ap. J. 472, 760–782.
Bonal, L., Quirico, E., Bourot-Denise, M. and Montagnac, G. (2006) Determination of the
petrologic type of CV3 chondrites by Raman spectroscopy of included organic matter.
Geochim. Cosmochim. Acta 70, 1849.
Bonal, L., Quirico, E., Flandinet, L. and Montagnac, G. (2016) Thermal history of type 3
chondrites from the Antarctic meteorite collection determined by Raman spectroscopy of their
polyaromatic carbonaceous matter. Geochim. Cosmochim. Acta 189, 312-337.
Bose, M., Floss, C., Stadermann, F.J., Stroud, R.M. and Speck, A.K. (2012) Circumstellar and
interstellar material in the CO3 chondrite ALHA77307: An isotopic and elemental
investigation. Geochim. Cosmochim. Acta 93, 77–101.
29
Bose, M., Zega, T.J. and Williams, P. (2014) Assessment of alteration processes on circumstellar
and interstellar grains in Queen Alexandra Range 97416. Earth Planet. Sci. Lett. 399, 128–
138.
Brearley, A.J. (1993) Matrix and fine-grained rims in the unequilibrated CO3 chondrite, ALHA
77307: Origins and evidence for diverse, primitive nebular dust components. Geochim.
Cosmochim. Acta 57, 1521–1550.
Burton, A.S., Elsila, J.E., Callahan, M.P., Martin, M.G., Glavin, D.P., Johnson, N.M. and Dworkin,
J.P. (2012) A propensity for n-omega-amino acids in thermally altered Antarctic meteorites.
Meteorit. Planet. Sci. 47, 374–386.
Busemann, H., Alexander, C.M.O'D. and Nittler, L.R. (2007) Characterization of insoluble
organic matter in meteorites by Raman spectroscopy. Meteorit. Planet. Sci. 42, 1387–1416.
Busemann, H., Nguyen, A.N., Cody, G.D., Hoppe, P., Kilcoyne, A.L.D., Stroud, R.M., Zega, T.J.
and Nittler, L.R. (2009) Ultra-primitive interplanetary dust particles from the comet
26P/Grigg-Skjellerup dust stream collection. Earth Planet. Sci. Lett. 288, 44–57.
Busemann, H., Young, A.F., Alexander, C.M.O'D., Hoppe, P., Mukhopadhyay, S. and Nittler,
L.R. (2006) Interstellar chemistry recorded in organic matter from primitive meteorites.
Science 312, 727–730.
Choi, B.-G., Huss, G.R. and Wasserburg, G.J. (1998) Presolar corundum and spinel in ordinary
chondrites: Origins from AGB stars and a supernova. Science 282, 1282–1289.
Ciesla, F.J. and Sandford, S.A. (2012) Organic synthesis via irradiation and warming of ice grains
in the solar nebula. Science 336, 452–454.
Cody, G.D., Alexander, C.M.O'D., Yabuta, H., Kilcoyne, A.L.D., Araki, T., Ade, H., Dera, P.,
Fogel, M., Militzer, B. and Mysen, B.O. (2008) Organic thermometry for chondritic parent
bodies. Earth Planet. Sci. Lett. 272, 446–455.
Cody, G.D., Heying, E., Alexander, C.M.O'D., Nittler, L.R., Kilcoyne, A.L.D., Sandford, S.A.
and Stroud, R.M. (2011) Establishing a molecular relationship between chondritic and
cometary organic solids. Proc. Nat. Acad. Sci. USA 108, 19171–19176
Dai, Z.R., Bradley, J.P., Joswiak, D.J., Brownlee, D.E., Hill, H.G.M. and Genge, M.J. (2002)
Possible in situ formation of meteoritic nanodiamonds in the early Solar System. Nature 418,
157–159.
Daulton, T.L., Bernatowicz, T.J., Lewis, R.S., Messenger, S., Stadermann, F.J. and Amari, S.
(2003) Polytype distribution of circumstellar silicon carbide - microstructural characterization
by transmission electron microscopy. Geochim. Cosmochim. Acta 67, 4743–4767.
Davidson, J., Busemann, H. and Franchi, I.A. (2012) A NanoSIMS and Raman spectroscopic
comparison of interplanetary dust particles from comet Grigg-Skjellerup and non-Grigg
Skjellerup collections. Meteorit. Planet. Sci. 47, 1748–1771.
Davidson, J., Busemann, H., Nittler, L.R., Alexander, C.M.O'D., Orthous-Daunay, F.-R., Franchi,
I.A. and Hoppe, P. (2014a) Abundances of presolar silicon carbide grains in primitive
meteorites determined by NanoSIMS. Geochim. Cosmochim. Acta 139, 248–266.
Davidson, J., Nittler, L.R., Alexander, C.M.O'D. and Stroud, R.M. (2014b) Petrography of very
primitive CO3 chondrites: Dominion Range 08006, Miller Range 07687, and four others.
Lunar Planet. Sci. XLV, #1384 (abstr.).
De Gregorio, B.T., Stroud, R.M., Nittler, L.R., Alexander, C.M.O'D., Bassim, N.D., Cody, G.D.,
Kilcoyne, A.L.D., Sandford, S.A., Milam, S.N., Nuevo, M. and Zega, T.J. (2013) Isotopic and
chemical variation of organic nanoglobules in primitive meteorites. Meteorit. Planet. Sci. 48,
804–828.
30
De Gregorio, B.T., Stroud, R.M., Nittler, L.R., Alexander, C.M.O'D., Kilcoyne, A.L.D. and Zega,
T.J. (2010) Isotopic anomalies in organic nanoglobules from Comet 81P/Wild 2: Comparison
to Murchison nanoglobules and isotopic anomalies induced in terrestrial organics by electron
irradiation. Geochim. Cosmochim. Acta 74, 4454–4470.
Dobrică, E., Engrand, C., Duprat, J., Gounelle, M., Leroux, H., Quirico, E. and Rouzaud, J.-N.
(2009) Connection between micrometeorites and Wild 2 particles: From Antarctic snow to
cometary ices. Meteorit. Planet. Sci. 44, 1643–1661.
Ehrenfreund, P. and Charnley, S.B. (2000) Organic molecules in the interstellar medium, comet
and meteorites: A voyage from dark clouds to the early Earth. Astron. Astrophys. 38, 427–483.
Floss, C. and Haenecour, P. (2016a) Meteorite Hills (MET) 00526: An unequilibrated ordinary
chondrite with high presolar grain abundances. Lunar Planet. Sci. XLVII, #1030 (abstr.).
Floss, C. and Haenecour, P. (2016b) Presolar silicate grains: Abundances, isotopic and elemental
compositions, and the effects of secondary processing. Geochemical Journal 50, 3-25.
Floss, C., Le Guillou, C. and Brearley, A. (2014) Coordinated NanoSIMS and FIB-TEM analyses
of organic matter and associated matrix materials in CR3 chondrites. Geochim. Cosmochim.
Acta 139, 1–25.
Floss, C. and Stadermann, F. (2009a) Auger Nanoprobe analysis of presolar ferromagnesian
silicate grains from primitive CR chondrites QUE 99177 and MET 00426. Geochim.
Cosmochim. Acta 73, 2415–2440.
Floss, C. and Stadermann, F.J. (2009b) High abundances of circumstellar and interstellar C-
anomalous phases in the primitive CR3 chondrites QUE 99177 and MET 00426. Ap. J. 697,
1242–1255.
Floss, C. and Stadermann, F.J. (2012) Presolar silicate and oxide abundances and compositions in
the ungrouped carbonaceous chondrite Adelaide and the K chondrite Kakangari: The effects
of secondary processing. Meteorit. Planet. Sci. 47, 992–1009.
Floss, C., Stadermann, F.J., Bradley, J., Dai, Z.R., Bajt, S. and Graham, G. (2004) Carbon and
nitrogen isotopic anomalies in an anhydrous interplanetary dust particle. Science 303, 1355–
1358.
Floss, C., Stadermann, F.J., Bradley, J.P., Dai, Z.R., Bajt, S., Graham, G. and Lea, A.S. (2006)
Identification of isotopically primitive interplanetary dust particles: A NanoSIMS isotopic
imaging study. Geochim. Cosmochim. Acta 70, 2371.
Gail, H.-P. (2010) Formation and evolution of minerals in accretion disks and stellar outflows, in:
Henning, T. (Ed.), Astromineralogy. Springer Berlin Heidelberg, Berlin, Heidelberg, pp. 61-
141.
Gail, H.-P. and Sedlmayr, E. (1999) Mineral formation in stellar winds. I. Condensation sequence
of silicate and iron grains in stationary oxygen rich outflows. Astron. Astrophys. 347, 594–616.
Garvie, L.A.J. and Buseck, P.R. (2004) Nanosized carbon-rich grains in carbonaceous chondrite
meteorites. Earth Planet. Sci. Lett. 224, 431–439.
Gehrels, N. (1986) Confidence limits for small numbers of events in astrophysical data. Ap. J. 303,
336–346.
Gourier, D., Robert, F., Delpoux, O., Binet, L., Vezin, H., Moissette, A. and Derenne, S. (2008)
Extreme deuterium enrichment of organic radicals in the Orgueil meteorite: Revisiting the
interstellar interpretation? Geochim. Cosmochim. Acta 72, 1914–1923.
Greshake, A. (1997) The primitive matrix components of the unique carbonaceous chondrite Acfer
094; A TEM study. Geochim. Cosmochim. Acta 61, 437–452.
31
Grossman, J.N. and Brearley, A.J. (2005) The onset of metamorphism in ordinary and
carbonaceous chondrites. Meteorit. Planet. Sci. 40, 87-122.
Grossman, L. and Steele, I.M. (1976) Amoeboid olivine aggregates in the Allende meteorite.
Gyngard, F., Nittler, L., Zinner, E. and Jose, J. (2010) Oxygen rich stardust grains from novae,
Geochim. Cosmochim. Acta 40, 149-155.
Nuclei in the Cosmos, p. 141.
Haenecour, P., Floss, C., José, J., Amari, S., Lodders, K., Jadhav, M., Wang, A. and Gyngard, F.
(2016) Coordinated analysis of two graphite grains from the CO3.0 LAP 031117 meteorite:
First identification of a CO nova graphite and a presolar iron sulfide subgrain. Ap. J. 825: 88.
Haenecour, P., Floss, C., Zega, T.J., Croat, T.K., Wang, A., Jolliff, B.L. and Carpenter, P. (2018)
Presolar silicates in the matrix and fine-grained rims around chondrules in primitive CO3.0
chondrites: Evidence for pre-accretionary aqueous alteration of the rims in the solar nebula.
Geochim. Cosmochim. Acta 221, 379-405.
Harju, E.R., Rubin, A.E., Ahn, I., Choi, B.-G., Ziegler, K. and Wasson, J.T. (2014) Progressive
aqueous alteration of CR carbonaceous chondrites. Geochim. Cosmochim. Acta 139, 267–292.
Hashizume, K., Chaussidon, M., Marty, B. and Terada, K. (2004) Protosolar carbon isotopic
composition: Implications for the origin of meteoritic organics. Ap. J. 600, 480–484.
Heck, P.R., Stadermann, F.J., Isheim, D., Auciello, O., Daulton, T.L., Davis, A.M., Elam, J.W.,
Floss, C., Hiller, J., Larson, D.J., Lewis, J.B., Mane, A., Pellin, M.J., Savina, M.R., Seidman,
D.N. and Stephan, T. (2014) Atom-probe analyses of nanodiamonds from Allende. Meteorit.
Planet. Sci. 49, 453–467.
Herd, C.D.K., Blinova, A., Simkus, D.N., Huang, Y., Tarozo, R., Alexander, C.M.O'D., Gyngard,
F., Nittler, L.R., Cody, G.D., Fogel, M.L., Kebukawa, Y., Kilcoyne, A.L.D., Hilts, R.W.,
Slater, G.F., Glavin, D.P., Dworkin, J.P., Callahan, M.P., Elsila, J.E., De Gregorio, B.T. and
Stroud, R.M. (2011) Origin and evolution of prebiotic organic matter as inferred from the
Tagish Lake meteorite. Science 332, 1304–1307.
Höfner, S., Grün, E. and Steinacker, J. (2009) Dust formation and winds around evolved stars: The
good, the bad and the ugly cases, in: Henning, T., Eberhard Grün, and Jürgen Steinacker (Eds.),
Cosmic Dust - Near and Far, ASP Conference Series, Vol. 414, San Francisco: Astromomical
Society of the Pacific, 2009, pp. 3-21.
Hoppe, P., Amari, S., Zinner, E., Ireland, T. and Lewis, R.S. (1994) Carbon, nitrogen, magnesium,
silicon, and titanium isotopic compositions of single interstellar silicon carbide grains from the
Murchison carbonaceous chondrite. Ap. J. 430, 870–890.
Hoppe, P., Leitner, J. and Kodolányi, J. (2015) New constraints on the abundances of silicate and
oxide stardust from supernovae in the Acfer 094 meteorite. Ap. J. 808.
Huss, G.R., Fahey, A.J., Gallino, R. and Wasserburg, G.J. (1994) Oxygen isotopes in circumstellar
Al2O3 grains from meteorites and stellar nucleosynthesis. Ap. J. 430, L81–84.
Huss, G.R. and Lewis, R.S. (1995) Presolar diamond, SiC, and graphite in primitive chondrites:
Abundances as a function of meteorite class and petrologic type. Geochim. Cosmochim. Acta
59, 115–160.
Huss, G.R., Meshik, A.P., Smith, J.B. and Hohenberg, C.M. (2003) Presolar diamond, silicon
carbide, and graphite in carbonaceous chondrites: implications for thermal processing in the
solar nebula. Geochim. Cosmochim. Acta 67, 4823–4848.
Joswiak, D.J., Brownlee, D.E. and Matrajt, G. (2013) First occurrence of a probable amoeboid
olivine aggregate in a "cometary" interplanetary dust particle. Lunar Planet. Sci. XLIV, #2410
(abstr.).
32
Keller, L.P., Messenger, S., Flynn, G.J., Clemett, S., Wirick, S. and Jacobsen, C. (2004) The nature
of molecular cloud material in interplanetary dust. Geochim. Cosmochim. Acta 68, 2577-2589.
Langer, W.D., Graedel, T.E., Frerking, M.A. and Armentrout, P.B. (1984) Carbon and oxygen
isotope fractionation in dense interstellar clouds. Ap. J. 277, 581-590.
Le Guillou, C., Bernard, S., Brearley, A.J. and Remusat, L. (2014) Evolution of organic matter in
Orgueil, Murchison and Renazzo during parent body aqueous alteration: In situ investigations.
Geochim. Cosmochim. Acta 131, 368–392.
Le Guillou, C. and Brearley, A. (2014) Relationships between organics, water and early stages of
aqueous alteration in the pristine CR3.0 chondrite MET 00426. Geochim. Cosmochim. Acta
131, 344–367.
Leitner, J., Hoppe, P., Floss, C., Hillion, F. and Henkel, T. (2018) Correlated nanoscale
characterization of a unique complex oxygen-rich stardust grain: Implications for circumstellar
dust formation. Geochim. Cosmochim. Acta 221, 255-274.
Leitner, J., Kodolányi, J., Hoppe, P. and Floss, C. (2012a) Laboratory analysis of presolar silicate
stardust from a nova. Ap. J. 754, L41.
Leitner, J., Vollmer, C., Floss, C., Zipfel, J. and Hoppe, P. (2016) Ancient stardust in fine-grained
chondrule dust rims from carbonaceous chondrites. Earth Planet. Sci. Lett. 434, 117–128.
Leitner, J., Vollmer, C., Hoppe, P. and Zipfel, J. (2012b) Characterization of presolar material in
the CR chondrite Northwest Africa 852. Ap. J. 745, 38.
Lugaro, M., Davis, A.M., Gallino, R., Pellin, M.J., Straniero, O. and Käppeler, F. (2003) Isotopic
compositions of strontium, zirconium, molybdenum, and barium in single presolar SiC grains
and asymptotic giant branch stars. Ap. J. 593, 486–508.
Lugaro, M., Karakas, A.I., Bruno, C.G., Aliotta, M., Nittler, L.R., Bemmerer, D., Best, A.,
Boeltzig, A., Broggini, C., Caciolli, A., Cavanna, F., Ciani, G.F., Corvisiero, P., Davinson, T.,
Depalo, R., Di Leva, A., Elekes, Z., Ferraro, F., Formicola, A., Fülöp, Z., Gervino, G.,
Guglielmetti, A., Gustavino, C., Gyürky, G., Imbriani, G., Junker, M., Menegazzo, R., Mossa,
V., Pantaleo, F.R., Piatti, D., Prati, P., Scott, D.A., Straniero, O., Strieder, F., Szücs, T., Takács,
M.P. and Trezzi, D. (2017) Origin of meteoritic stardust unveiled by a revised proton-capture
rate of 17O. Nature Astronomy 1, 0027.
Lyons, J.R., Gharib-Nezhad, E. and Ayres, T.R. (2017) The carbon isotope composition of the
Sun. Lunar Planet. Sci. XLVIII, #2309 (abstr.).
Marty, B., Chaussidon, M., Wiens, R.C., Jurewicz, A.J.G. and Burnett, D.S. (2011) A 15N-poor
isotopic composition for the Solar System as shown by Genesis solar wind samples. Science
332, 1533.
Matrajt, G., Messenger, S., Brownlee, D. and Joswiak, D. (2012) Diverse forms of primordial
organic matter identified in interplanetary dust particles. Meteorit. Planet. Sci. 47, 525–549.
Matsuura, M., Bernard-Salas, J., Lloyd Evans, T., Volk, K.M., Hrivnak, B.J., Sloan, G.C., Chu,
Y.-H., Gruendl, R., Kraemer, K.E., Peeters, E., Szczerba, R., Wood, P.R., Zijlstra, A.A., Hony,
S., Ita, Y., Kamath, D., Lagadec, E., Parker, Q.A., Reid, W.A., Shimonishi, T., Van Winckel,
H., Woods, P.M., Kemper, F., Meixner, M., Otsuka, M., Sahai, R., Sargent, B.A., Hora, J.L.
and McDonald, I. (2014) Spitzer Space Telescope spectra of post-AGB stars in the Large
Magellanic Cloud – polycyclic aromatic hydrocarbons at low metallicities. MNRAS 439, 1472-
1493.
McKeegan, K.D., Aleon, J., Bradley, J., Brownlee, D., Busemann, H., Butterworth, A.,
Chaussidon, M., Fallon, S., Floss, C., Gilmour, J., Gounelle, M., Graham, G., Guan, Y., Heck,
P.R., Hoppe, P., Hutcheon, I.D., Huth, J., Ishii, H., Ito, M., Jacobsen, S.B., Kearsley, A.,
33
Leshin, L.A., Liu, M.-C., Lyon, I., Marhas, K., Marty, B., Matrajt, G., Meibom, A., Messenger,
S., Mostefaoui, S., Mukhopadhyay, S., Nakamura-Messenger, K., Nittler, L., Palma, R., Pepin,
R.O., Papanastassiou, D.A., Robert, F., Schlutter, D., Snead, C.J., Stadermann, F.J., Stroud,
R., Tsou, P., Westphal, A., Young, E.D., Ziegler, K., Zimmermann, L. and Zinner, E. (2006)
Isotopic compositions of cometary matter returned by Stardust. Science 314, 1724–1728.
Messenger, S. (2000) Identification of molecular cloud material in interplanetary dust particles.
Messenger, S., Keller, L.P. and Lauretta, D.S. (2005) Supernova Olivine from Cometary Dust.
Nature 404, 968–971.
Science 309, 737–741.
Molster F., Waters L. and Kemper F. (2010) The Mineralogy of Interstellar and Circumstellar Dust
in Galaxies. In: Henning T. (Ed.) Astromineralogy. Lecture Notes in Physics, vol 815.
Springer, Berlin, Heidelberg, pp. 142-201.
Nakamura-Messenger, K., Messenger, S., Keller, L.P., Clemett, S.J. and Zolensky, M.E. (2006)
Organic globules in the Tagish Lake meteorite: Remnants of the protosolar disk. Science 314,
1439–1442.
Nguyen, A.N., Nakamura-Messenger, K., Messenger, S., Keller, L.P. and Klöck, W. (2014)
Identification of a compound spinel and silicate presolar grain in a chondritic interplanetary
dust particle. Lunar Planet. Sci. XLV, #2351 (abstr.).
Nguyen, A.N., Nittler, L.R., Stadermann, F.J., Stroud, R.M. and Alexander, C.M.O'D. (2010)
Coordinated analyses of presolar grains in the Allan Hills 77307 and Queen Elizabeth Range
99177 meteorites. Ap. J. 719, 166–189.
Nguyen, A.N., Stadermann, F.J., Zinner, E., Stroud, R.M., Alexander, C.M.O'D. and Nittler, L.R.
(2007) Characterization of presolar silicate and oxide grains in primitive carbonaceous
chondrites. Ap. J. 656, 1223–1240.
Nittler, L., Alexander, C.M.O'D., Gao, X., Walker, R.M. and Zinner, E. (1994) Interstellar oxide
grains from the Tieschitz ordinary chondrite. Nature 370, 443–446.
Nittler, L.R. (2009) On the mass and metallicity distributions of the parent AGB stars of O-rich
presolar stardust grains. Pub. Astron. Soc. Aust. 26, 271–277.
Nittler, L.R., Alexander, C.M.O'D., Gallino, R., Hoppe, P., Nguyen, A., Stadermann, F. and
Zinner, E.K. (2008) Aluminum-, calcium- and titanium-rich oxide stardust in ordinary
chondrite meteorites. Ap. J. 682, 1450–1478.
Nittler, L.R., Alexander, C.M.O'D., Gao, X., Walker, R.M. and Zinner, E. (1997) Stellar
sapphires: The properties and origins of presolar Al2O3 in meteorites. Ap. J. 483, 475–495.
Nittler, L.R. and Ciesla, F. (2016) Astrophysics with extraterrestrial materials. Ann. Rev. Astron.
Astrophys. 54, 53–93.
Nollett, K.M., Busso, M. and Wasserburg, G.J. (2003) Cool bottom processes on the thermally-
pulsing AGB and the isotopic composition of circumstellar dust grains. Ap. J. 582, 1036–1058.
Nuth, J.A., III, Johnson, N.M. and Manning, S. (2008) A self-perpetuating catalyst for the
production of complex organic molecules in protostellar nebulae. Ap. J. 673, L225–L228.
Palmerini, S., Trippella, O. and Busso, M. (2017) A deep mixing solution to the aluminum and
oxygen Isotope puzzles in presolar grains. MNRAS 467, 1193-1201.
Peeters, Z., Changela, H.G., Stroud, R.M., Alexander, C.M.O'D. and Nittler, L.R. (2012)
Coordinated analysis of in situ organic material in the CR chondrite QUE 99177. Lunar Planet.
Sci. XLIII, #1659 (abstr.).
Pendleton, Y.J., Sandford, S.A., Allamandola, L.J., Tielens, A.G.G.M. and Sellgren, K. (1994)
Near-infrared absorption spectroscopy of interstellar hydrocarbon grains. Ap. J. 437, 683–696.
34
Quirico, E., Raynal, P.I. and Bourot-Denise, M. (2003) Metamorphic grade of organic matter in
six unequilibrated ordinary chondrites. Meteorit. Planet. Sci. 38, 795–811.
Remusat, L., Guan, Y., Wang, Y. and Eiler, J.M. (2010) Accretion and preservation of D-rich
organic particles in carbonaceous chondrites: Evidence for important transport in the early
solar system nebula. Ap. J. 713, 1048–1058.
Robert, F. and Epstein, S. (1982) The concentration and isotopic composition of hydrogen, carbon
and nitrogen in carbonaceous meteorites. Geochim. Cosmochim. Acta 46, 81–95.
Sandford, S.A., Bernstein, M.P. and Dworkin, J.P. (2001) Assessment of the interstellar processes
leading to deuterium enrichment in meteoritic organics. Meteorit. Planet. Sci. 36, 1117–1133.
Scott, E.R.D. and Krot, A.N. (2003) Chondrites and their Components, in: Davis, A.M. (Ed.),
Meteorites, Comets and Planets (Vol. 1), Treatise on Geochemistry (eds: H. D. Holland and
K. K. Turekian). Elsevier-Pergamon, Oxford, pp. 143-200.
Slodzian, G., Hillion, F., Stadermann, F.J. and Zinner, E. (2004) QSA influences on isotopic ratio
measurements. Applied Surface Science 231–232, 874–877.
Stadermann, F.J., Floss, C., Bose, M. and Lea, A.S. (2009) The use of Auger spectroscopy for the
in situ elemental characterization of sub-micrometer presolar grains. Meteorit. Planet. Sci. 44,
1033–1049.
Stroud, R.M., Chisholm, M.F., Heck, P.R., Alexander, C.M.O'D. and Nittler, L.R. (2011)
Supernova shock-wave-induced co-formation of glassy carbon and nanodiamond. Ap. J. 738.
Stroud, R.M., De Gregorio, B.T., Nittler, L.R. and Alexander, C.M.O'D. (2014) Comparative
transmission electron microscopy studies of presolar silicate and oxide grains from the
Dominion Range 08006 and Northwest Africa 5958 meteorites. Lunar Planet. Sci. XLV, #2806
(abstr.).
Stroud, R.M., Nittler, L.R. and Alexander, C.M.O'D. (2004) Polymorphism in presolar Al2O3
grains from asymptotic giant branch stars. Science 305, 1455–1457.
Stroud, R.M., Nittler, L.R. and Alexander, C.M.O'D. (2013) Analytical electron microscopy of a
CAI-like presolar grain and associated fine-grained matrix materials in the Dominion Range
08006 CO3 meteorite. Lunar Planet. Sci. XLIV, #2315 (abstr.).
Takigawa, A., Stroud, R.M., Nittler, L.R., Vicenzi, E.P., Herzing, A., Alexander, C.M.O'D. and
Huss, G.R. (2014) Crystal structure, morphology, and isotopic compositions of presolar Al2O3
grains in unequilibrated ordinary chondrites. Lunar Planet. Sci. XLV, #1465 (abstr.).
Vollmer, C., Brenker, F.E., Hoppe, P. and Stroud, R.M. (2009a) Direct laboratory analysis of
silicate stardust from red giant stars. Ap. J. 700, 774–782.
Vollmer, C., Hoppe, P. and Brenker, F.E. (2013) Transmission electron microscopy of Al-rich
silicate stardust from asymptotic giant branch stars. Ap. J. 769, 61.
Vollmer, C., Hoppe, P., Brenker, F.E. and Holzapfel, C. (2007) Stellar MgSiO3 perovskite: A
shock-transformed stardust silicate found in a meteorite. Ap. J. 666, L49–L52.
Vollmer, C., Hoppe, P., Stadermann, F.J., Floss, C. and Brenker, F.E. (2009b) NanoSIMS analysis
and Auger electron spectroscopy of silicate and oxide stardust from the carbonaceous chondrite
Acfer 094. Geochim. Cosmochim. Acta 73, 7127–7149.
Vollmer, C., Kepaptsoglou, D., Leitner, J., Busemann, H., Spring, N.H., Ramasse, Q.M., Hoppe,
P. and Nittler, L.R. (2014) Fluid-induced organic synthesis in the solar nebula recorded in
extraterrestrial dust from meteorites. Proc. Nat. Acad. Sci. USA 111, 15338–15343.
Wasserburg, G.J., Boothroyd, A.I. and Sackmann, I.-J. (1995) Deep circulation in red giant stars:
A solution to the carbon and oxygen isotope puzzles? Ap. J. 447, L37–40.
35
Acta 47, 2199–2216.
Zega, T.J., Alexander, C.M.O'D., Nittler, L.R. and Stroud, R.M. (2011) A transmission electron
microscopy study of presolar hibonite. Ap. J. 730, 83–92.
Zega, T.J., Nittler, L.R., Gyngard, F., Alexander, C.M.O'D., Stroud, R.M. and Zinner, E.K. (2014)
A transmission electron microscopy study of presolar spinel. Geochim. Cosmochim. Acta 124,
152–169.
Zhao, X., Floss, C., Lin, Y. and Bose, M. (2013) Stardust investigation into the CR chondrite
Yada, T., Floss, C., Stadermann, F.J., Zinner, E., Nakamura, T., Noguchi, T. and Lea, A.S. (2008)
Stardust in Antarctic micrometeorites. Meteorit. Planet. Sci. 43, 1287–1298.
Yang, J. and Epstein, S. (1983) Interstellar organic matter in meteorites. Geochim. Cosmochim.
Grove Mountain 021710. Ap. J. 769.
Zinner, E. (2014) 1.4 - Presolar grains, in: Davis, A.M. (Ed.), Meteorites and Cosmochemical
Processes (Vol. 1), Treatise on Geochemistry (Second Edition, eds: H. D. Holland and K. K.
Turekian). Elsevier-Pergamon, Oxford, pp. 181–213.
Zinner, E., Amari, S., Guinness, R., Nguyen, A., Stadermann, F.J., Walker, R.M. and Lewis, R.S.
(2003) Presolar spinel grains from the Murray and Murchison carbonaceous chondrites.
Geochim. Cosmochim. Acta 67, 5083–5095.
36
Table 1 Areas analyzed by NanoSIMS, resultant number of presolar grains and matrix-normalized
abundances of O-rich presolar grains.
# Presolar Grains
Abundance of presolar O-rich grains (ppm)
Analyzed Area
(µm2)
10,820
7,500
3,000
6,550
Area 2
Area 3
Area 6
Area 7
Total
O-rich
43
36
11
11
101
SiC
7
5
0
5
17
Average
299
213
217
155
257a
Error –
(1σ)
Error +
(1σ)
Error –
(2σ)
41
44
76
Error +
(2σ)
96
a Excluding Area 7; errors estimated from Monte Carlo calculations (see text)
37
Table 2. Sizes and isotopic compositions of presolar O-rich grains.
Size
(nm)
207
237
545
229
192
261
275
522
241
333
938
382
257
282
268
330
363
324
225
176
369
245
268
361
253
257
245
245
350
176
384
318
427
268
257
275
330
Grain
DOM-1
DOM-2
DOM-3
DOM-4
DOM-5
DOM-6
DOM-7
DOM-8
DOM-9
DOM-10
DOM-11
DOM-12
DOM-13
DOM-14
DOM-15
DOM-16
DOM-17
DOM-18
DOM-19
DOM-20
DOM-21
DOM-22
DOM-23
DOM-24
DOM-25
DOM-26
DOM-27
DOM-28
DOM-29
DOM-30
DOM-31
DOM-32
DOM-33
DOM-34
DOM-35
DOM-36
DOM-37
Size (nm,
corrected) Group
17O/16O
(× 10-4)
18O/16O
(× 10-3)
Si-/O-
(× 10-3)
168
205
532
195
150
232
248
508
209
310
930
362
227
255
240
307
343
301
190
129
349
214
240
340
223
227
214
214
329
129
365
294
410
240
227
248
307
1
1
1
1
3
1
1
1
1
1
1
1
1
1
1
1
1
1
1
4
1
1
2
1
1
1
1
1
4
1
1
1
1
1
1
1
1
1.990 ± 0.049
1.944 ± 0.046
1.820 ± 0.023
1.892 ± 0.056
1.700 ± 0.057
2.010 ± 0.050
1.348 ± 0.055
1.972 ± 0.025
1.978 ± 0.066
1.949 ± 0.051
1.917 ± 0.034
1.576 ± 0.029
2.006 ± 0.039
1.994 ± 0.046
2.043 ± 0.041
1.473 ± 0.031
2.050 ± 0.051
1.325 ± 0.035
1.886 ± 0.060
2.259 ± 0.088
1.989 ± 0.057
1.993 ± 0.053
1.048 ± 0.063
1.944 ± 0.028
1.967 ± 0.039
1.604 ± 0.077
2.092 ± 0.045
1.891 ± 0.043
2.490 ± 0.047
2.573 ± 0.123
1.866 ± 0.032
2.001 ± 0.039
1.724 ± 0.037
1.636 ± 0.046
2.066 ± 0.046
1.918 ± 0.050
1.966 ± 0.048
8.81 ± 0.10
13.50 ± 0.12
10.21 ± 0.05
7.66 ± 0.11
8.81 ± 0.12
7.38 ± 0.10
10.14 ± 0.13
10.19 ± 0.06
9.19 ± 0.14
5.63 ± 0.09
7.74 ± 0.07
16.89 ± 0.09
13.88 ± 0.10
11.96 ± 0.11
13.05 ± 0.11
12.21 ± 0.08
17.57 ± 0.15
15.20 ± 0.10
18.11 ± 0.18
15.90 ± 0.24
15.11 ± 0.16
12.65 ± 0.13
12.27 ± 0.16
8.88 ± 0.06
10.62 ± 0.09
15.39 ± 0.22
11.93 ± 0.11
13.33 ± 0.11
9.33 ± 0.09
5.11 ± 0.18
10.57 ± 0.07
8.06 ± 0.08
4.60 ± 0.06
7.32 ± 0.09
4.84 ± 0.07
8.55 ± 0.11
9.59 ± 0.11
5.30 ± 0.26
5.26 ± 0.24
9.91 ± 0.18
5.49 ± 0.29
3.42 ± 0.26
5.56 ± 0.28
6.09 ± 0.32
8.83 ± 0.17
6.34 ± 0.39
6.14 ± 0.30
4.97 ± 0.18
6.21 ± 0.17
7.05 ± 0.24
8.20 ± 0.31
5.03 ± 0.21
6.57 ± 0.19
23.63 ± 0.57
6.75 ± 0.21
5.89 ± 0.34
5.97 ± 0.48
7.47 ± 0.37
6.90 ± 0.33
5.76 ± 0.36
6.21 ± 0.16
5.54 ± 0.22
5.47 ± 0.42
7.17 ± 0.28
5.30 ± 0.23
3.72 ± 0.19
3.78 ± 0.46
8.12 ± 0.20
8.60 ± 0.25
9.11 ± 0.24
5.37 ± 0.23
4.92 ± 0.22
6.29 ± 0.27
7.08 ± 0.28
38
AlO-/O-
(× 10-4)
4.5 ± 0.2
4.4 ± 0.2
13.2 ± 0.2
10.9 ± 0.4
2.0 ± 0.2
25.6 ± 0.6
5.1 ± 0.3
23.2 ± 0.3
10.0 ± 0.5
41.8 ± 0.7
21.7 ± 0.4
13.6 ± 0.2
15.1 ± 0.3
15.0 ± 0.4
16.7 ± 0.4
28.2 ± 0.4
7.8 ± 0.3
24.6 ± 0.4
8.4 ± 0.4
7.2 ± 0.5
3.5 ± 0.2
8.4 ± 0.3
42.7 ± 0.9
120.5 ± 0.7
180.7 ± 1.2
15.9 ± 0.7
37.4 ± 0.6
73.1 ± 0.8
5.2 ± 0.2
27.0 ± 1.3
15.3 ± 0.3
22.3 ± 0.4
108.6 ± 0.9
12.5 ± 0.4
46.1 ± 0.7
24.3 ± 0.6
6.4 ± 0.3
248
252
240
281
214
223
209
255
195
227
420
185
223
270
274
185
270
156
129
168
205
190
162
323
223
323
094
185
240
094
267
156
274
244
156
223
227
185
223
349
236
275
279
268
305
245
253
241
282
229
257
436
220
253
296
299
220
296
197
176
207
237
225
202
344
253
344
153
220
268
153
292
197
299
272
197
253
257
220
253
369
264
DOM-38
DOM-39
DOM-40
DOM-41
DOM-42
DOM-43
DOM-44
DOM-45
DOM-46
DOM-47
DOM-48
DOM-49
DOM-50
DOM-51
DOM-52
DOM-53
DOM-54
DOM-55
DOM-56
DOM-57
DOM-58
DOM-59
DOM-60
DOM-61
DOM-62
DOM-63
DOM-64
DOM-65
DOM-66
DOM-67
DOM-68
DOM-69
DOM-70
DOM-71
DOM-72
DOM-73
DOM-74
DOM-75
DOM-76
DOM-77
DOM-78
1
1
1
1
4
4
1
1
4
1
1
1
1
1
1
1
1
4
1
4
4
1
4
1
1
1
4
4
1
4
1
1
1
1
4
1
1
4
1
1
1
2.013 ± 0.044
1.868 ± 0.046
1.512 ± 0.057
2.034 ± 0.038
1.610 ± 0.064
2.365 ± 0.058
1.635 ± 0.073
1.654 ± 0.031
2.685 ± 0.053
1.576 ± 0.034
1.973 ± 0.022
2.052 ± 0.059
2.009 ± 0.045
1.657 ± 0.045
1.962 ± 0.034
1.946 ± 0.037
1.926 ± 0.036
2.449 ± 0.095
2.017 ± 0.080
2.488 ± 0.082
3.101 ± 0.146
1.964 ± 0.055
2.809 ± 0.119
2.042 ± 0.024
1.684 ± 0.036
1.964 ± 0.033
2.559 ± 0.111
2.664 ± 0.059
1.909 ± 0.039
2.248 ± 0.059
1.628 ± 0.031
1.974 ± 0.066
1.671 ± 0.039
1.910 ± 0.032
2.285 ± 0.051
2.096 ± 0.059
1.973 ± 0.037
2.709 ± 0.047
1.481 ± 0.038
1.259 ± 0.052
1.915 ± 0.060
7.37 ± 0.09
8.06 ± 0.10
9.45 ± 0.13
10.97 ± 0.10
9.92 ± 0.16
7.78 ± 0.11
12.14 ± 0.18
8.84 ± 0.06
16.53 ± 0.12
8.76 ± 0.07
12.80 ± 0.06
19.24 ± 0.15
12.67 ± 0.10
22.83 ± 0.12
17.33 ± 0.09
5.58 ± 0.06
17.33 ± 0.10
14.02 ± 0.23
7.12 ± 0.15
7.51 ± 0.14
11.88 ± 0.29
2.56 ± 0.06
1.47 ± 0.09
12.32 ± 0.06
10.53 ± 0.08
13.22 ± 0.09
12.59 ± 0.25
14.01 ± 0.14
8.16 ± 0.08
13.93 ± 0.15
16.49 ± 0.10
20.68 ± 0.23
17.33 ± 0.12
7.22 ± 0.06
14.38 ± 0.14
20.37 ± 0.20
11.20 ± 0.09
10.52 ± 0.09
13.91 ± 0.10
15.10 ± 0.08
13.57 ± 0.11
5.21 ± 0.22
6.97 ± 0.27
5.07 ± 0.28
6.25 ± 0.21
5.71 ± 0.33
3.76 ± 0.23
6.43 ± 0.42
6.18 ± 0.18
5.22 ± 0.24
6.16 ± 0.19
4.70 ± 0.11
5.10 ± 0.30
5.91 ± 0.25
4.44 ± 0.22
6.02 ± 0.19
5.13 ± 0.19
5.40 ± 0.19
4.36 ± 0.40
6.22 ± 0.45
4.30 ± 0.34
4.61 ± 0.57
5.34 ± 0.29
3.18 ± 0.44
4.73 ± 0.12
5.45 ± 0.19
7.58 ± 0.21
3.35 ± 0.44
4.05 ± 0.24
8.21 ± 0.26
4.12 ± 0.26
5.94 ± 0.17
7.70 ± 0.42
4.75 ± 0.20
6.08 ± 0.18
3.69 ± 0.21
6.39 ± 0.34
5.61 ± 0.20
3.86 ± 0.18
6.24 ± 0.22
6.01 ± 0.16
6.93 ± 0.32
39
34.5 ± 0.6
31.2 ± 0.6
9.8 ± 0.4
57.1 ± 0.7
11.8 ± 0.5
17.6 ± 0.5
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
n.m.
DOM-79
DOM-80
DOM-81
DOM-82
DOM-83
DOM-84
DOM-85
DOM-86
DOM-87
DOM-88
DOM-89
DOM-90
DOM-91
DOM-92
DOM-93
DOM-94
DOM-95
DOM-96
DOM-97
DOM-98
DOM-99
DOM-100
DOM-101
416
416
279
211
341
237
299
182
302
253
402
279
296
220
207
358
192
533
341
387
523
377
379
398
398
252
174
320
205
274
136
277
223
383
252
270
185
168
337
150
519
320
368
509
357
360
1
1
1
1
1
1
1
1
1
1
2
1
1
1
1
1
4
4
1
1
1
1
1
5.41 ± 0.13
8.17 ± 0.16
6.40 ± 0.20
5.93 ± 0.26
8.98 ± 0.20
5.37 ± 0.25
15.41 ± 0.83
4.72 ± 0.21
5.09 ± 0.15
5.13 ± 0.18
7.80 ± 0.18
5.65 ± 0.20
5.54 ± 0.20
4.58 ± 0.26
5.37 ± 0.29
5.25 ± 0.14
3.44 ± 0.31
3.51 ± 0.17
5.14 ± 0.18
5.57 ± 0.17
6.03 ± 0.14
6.61 ± 0.19
6.17 ± 0.22
2.067 ± 0.042
2.056 ± 0.026
1.460 ± 0.037
2.053 ± 0.050
1.810 ± 0.030
1.925 ± 0.050
2.082 ± 0.099
2.173 ± 0.046
1.951 ± 0.031
1.957 ± 0.037
1.124 ± 0.031
1.935 ± 0.039
1.725 ± 0.038
1.655 ± 0.056
2.105 ± 0.057
1.917 ± 0.028
2.368 ± 0.077
4.773 ± 0.060
1.952 ± 0.035
1.477 ± 0.032
1.891 ± 0.025
1.372 ± 0.034
1.619 ± 0.039
12.25 ± 0.06
5.02 ± 0.04
6.25 ± 0.06
5.02 ± 0.08
5.95 ± 0.05
5.96 ± 0.09
7.00 ± 0.18
2.79 ± 0.05
2.82 ± 0.04
3.78 ± 0.05
4.46 ± 0.05
4.99 ± 0.06
3.60 ± 0.05
2.82 ± 0.07
4.79 ± 0.09
2.93 ± 0.04
4.07 ± 0.11
3.30 ± 0.05
4.16 ± 0.05
5.68 ± 0.06
4.03 ± 0.04
4.21 ± 0.05
7.23 ± 0.08
n.m.
32.6 ± 0.3
26.1 ± 0.4
79.7 ± 1.0
37.8 ± 0.4
43.0 ± 0.7
51.4 ± 1.5
77.6 ± 0.9
35.6 ± 0.4
28.3 ± 0.4
223.5 ± 1.0
27.5 ± 0.5
49.4 ± 0.6
93.7 ± 1.2
73.6 ± 1.1
83.1 ± 0.6
48.6 ± 1.2
2.2 ± 0.1
158.9 ± 1.0
139.7 ± 0.9
38.5 ± 0.4
43.1 ± 0.5
23.2 ± 0.5
40
Table 3 Sizes and isotopic compositions of C-anomalous grains.
Grain
A3b2_4
A3_9
A3b2_4b
A3_11
A2_6
A2d_3
A2r2_20
A2r_3
A2r_16
A2r2_15
A2r_17
A7_16
A7_49
A7_59
A7_66
A7_41
A3cCN_7
A2c_20
A2r_3
A2r2_15
A3_14a
A3_14b
A3_20
A3_22
A3b2_10
A3b2_11
A3b2_16
A3b2_17
A3b2_29
A3b2_30
A3b2_5
A3cCN_1a
A3cCN_1b
A3cCN_2a
A3cCN_2b
A3cCN_2c
A3cCN_2d
Phase
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
SiC
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
Diameter
(nm)
117
205
94
263
214
263
209
168
129
121
368
354
252
388
420
413
252
195
113
291
277
185
232
236
190
252
317
263
259
304
294
236
370
263
291
334
403
δ15N
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
-73 ± 163
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
n. m.
-130 ± 42
268 ± 58
-115 ± 124
529 ± 74
213 ± 64
298 ± 73
δ13C
2020 ± 156
831 ± 46
336 ± 76
304 ± 29
305 ± 49
1026 ± 51
493 ± 109
514 ± 110
1034 ± 209
515 ± 142
368 ± 60
276 ± 27
348 ± 58
445 ± 48
208 ± 30
2240 ± 90
251 ± 42
-294 ± 45
-493 ± 111
-208 ± 23
130 ± 17
-203 ± 30
-185 ± 24
143 ± 24
-163 ± 26
-135 ± 22
-237 ± 25
130 ± 22
137 ± 20
200 ± 20
128 ± 17
3 ± 17
51 ± 17
-129 ± 38
18 ± 15
17 ± 15
28 ± 21
41
A3cCN_2e
A3cCN_3a
A3cCN_3b
A3cCN_3c
A3cCN_3d
A3cCN_3e
A3cCN_3f
A3cCN_4a
A3cCN_4b
A3cCN_4c
A3cCN_4d
A3cCN_5a
A3cCN_5b
A3cCN_5c
A3cCN_5d
A3cCN_6a
A3cCN_6b
A3cCN_6c
A3cCN_6d
A3cCN_6e
A3cCN_6f
A3cCN_6g
A3cCN_6h
A3cCN_6i
A3cCN_6j
A3cCN_7a
A3cCN_7b
A6b_5
A7_60
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
organic
nanoglobule
organic
organic
organic
270
252
281
240
252
291
291
310
310
248
263
301
190
252
259
298
255
368
291
307
301
223
415
136
255
360
343
340
502
-35 ± 38
44 ± 16
32 ± 13
-14 ± 25
17 ± 27
-19 ± 29
-54 ± 36
-37 ± 14
-7 ± 20
9 ± 24
-27 ± 24
-156 ± 46
10 ± 21
-12 ± 19
9 ± 41
-232 ± 13
42 ± 16
-196 ± 14
2 ± 16
-197 ± 17
-29 ± 29
134 ± 24
30 ± 33
-154 ± 45
-146 ± 41
98 ± 9
-8 ± 10
-290 ± 21
-300 ± 16
-294 ± 80
670 ± 67
496 ± 85
-248 ± 80
340 ± 93
-210 ± 65
449 ± 107
659 ± 77
-188 ± 45
871 ± 143
338 ± 93
184 ± 207
459 ± 77
-256 ± 56
-297 ± 93
-225 ± 45
342 ± 83
-215 ± 49
-101 ± 32
-183 ± 55
-234 ± 73
543 ± 69
337 ± 108
-47 ± 164
273 ± 146
361 ± 72
288 ± 66
n. m.
-264 ± 29
Where n. m. = not measured.
42
Table 4: Composition and structure of presolar grains analyzed by TEM
Grain
DOM-3
DOM-8
DOM-17
DOM-18
DOM-77
Composition
Mg0.45Fe0.74Ni0.07Ca0.03Al0.05Si0.82O3
Mg1.61Fe0.33Ni0.03Al0.02Si1.00O4
Mg0.33Fe0.86Ca0.03Al0.06Si0.84O3
Mg1.85Fe0.14Ni0.01Si1.00O4
Whole grain average Mg1.24Fe0.43Ni0.09
Al0.02Ca0.01Cr0.01Si1.03O4
Structure
Amorphous
Unknown (olivine?)
Amorphous
Olivine
Concentrically zoned, polycrystalline olivine
+ Ca-rich nanocrystal + Al-rich oxide
43
Figure 1. Reflected-light (center) and secondary electron (SE, top and bottom) images of DOM
08006 thin section. The yellow boxes and SE images indicate areas analyzed by NanoSIMS
(Table 1). NanoSIMS raster areas are clearly visible in the SE images.
44
Figure 2. Left: 17O/16O ratio image and line profile data for pixels indicated by dashed line for
presolar grain DOM-80. Data are smoothed with a 3-pixel boxcar, such that the ratio and error
bar for a given pixel are calculated from the total counts of that pixel and its two surrounding
ones. Right: corresponding "sigma" image (color scale is absolute value) and line profile,
where value in each pixel is difference (in standard deviations) of measured 17O/16O ratio
relative to terrestrial composition. Only pixels with values that lie within the full width at half
maximum (FWHM) of the isotopically anomalous region are included in definition of the
presolar grain ROI (grey area in profile plot and outline in inset).
45
Figure 3. Example SEM and NanoSIMS images of one 10 × 10 µm area of DOM 08006. (a)
SEM secondary electron image; (b) NanoSIMS 16O ion image; (c) NanoSIMS 12C image, (d)
SEM secondary electron image of 13C- and 15N-rich organic nanoglobule, area marked by box
in (a,f,g); (e) NanoSIMS δ17O/16O image with a 17O-rich presolar silicate (DOM-79, Table 2)
indicated; (f) NanoSIMS δ12C/13C image with a 13C-rich presolar SiC indicated; and (g)
NanoSIMS δ15N/14N image with two clear "hotspots" of enhanced 15N abundance, one of
which corresponds to the nanoglobule in panel (d). Unless otherwise noted, scale bars are 1
µm. Delta values are deviations from a standard ratio
in permil, defined as:
δR=(Rmeasured/Rstandard - 1) ×103.
46
Figure 4. Size distribution of 101 presolar O-rich grains identified in DOM 08006. Grey histogram
indicates measured equivalent diameters while the dotted histogram indicates data corrected
for 120-nm NanoSIMS beam broadening (see text for details).
47
Figure 5. a) Histogram of NanoSIMS 27Al16O-/16O- ion ratios for 65 presolar grains from DOM
08006; the maximum bin contains 8 grains. The dashed line indicates the threshold assumed
to differentiate between silicate and oxide grains. b) NanoSIMS 28Si-/16O- secondary ion ratios
plotted against AlO-/O- ion ratios for the same presolar grains. The grey box indicates grains
that are assumed to be Al-rich oxide grains, while grains with lower AlO-/O- ratios are assumed
to be silicates.
48
Figure 6. O-isotopic ratios in presolar O-rich grains from DOM 08006 compared to those
previously reported for presolar oxides and silicates (e.g., Nittler et al., 2008; Zinner, 2014;
Floss and Haenecour, 2016b). Ellipses indicate grain groups defined by Nittler et al. (1997).
49
Figure 7. Histograms of O isotopic ratios measured in presolar oxide and silicate grains (see Figure
6 for data sources). Top: Previously reported data for silicates and oxides are compared;
presolar silicates show a narrower distribution of 17O/16O and 18O/16O ratios than presolar oxide
grains, reflecting isotope dilution in in situ SIMS analyses. Bottom: Data for 101 DOM 08006
presolar grains compared to the previous silicate data (re-scaled to fit on plots). The DOM
08006 data span narrower ranges of O-isotope ratios, most likely reflecting both the limited
statistics as well as a conservative choice for identifying grains in ion images (see text).
50
Figure 8. C-isotopic ratios of DOM 08006 presolar SiC grains compared to literature data for
~16,000 isolated grains from acid residues (top panel; Zinner, 2014) and 118 grains found in
situ in previous studies (bottom panel; Floss and Stadermann, 2009a; Nguyen et al., 2010; Bose
et al., 2012; Leitner et al., 2012b; Haenecour et al., 2018). The vertical solid line indicates the
terrestrial ratio of 89 and the vertical dashed line indicates the peak of the isolated grain
distribution at 12C/13C=57.
51
Figure 9. The distribution of δ13C values in 25 C-anomalous grains in DOM 00086 (top) is
compared to those (bottom) measured in CR chondrites MET 00426 and QUE 99177 (Floss
and Stadermann, 2009b).
52
Figure 10. C- and N-isotopic ratios, expressed as δ-values, of organic grains that are anomalous in
one or both of these elements from DOM 08006 (present study), QUE 99177 and MET 00426
(Floss and Stadermann, 2009b), and ALH 77307 (Bose et al., 2012). Small grey circles are
DOM 08006 data that do not meet our criteria for being considered anomalous (see text). F04
indicates an anomalous grain found in an IDP (Floss et al., 2004), B06 indicates an isotopic
hotspot in insoluble organic matter from a CR chondrite (Busemann et al., 2006), and D10
indicates a carbonaceous nanoglobule from comet Wild-2 (De Gregorio et al., 2010). The
dotted lines indicate the terrestrial values. Dashed ellipse indicates population of 15N-enriched,
13C-depleted organic grains seen in several other extraterrestrial samples, but not DOM 08006.
Isotopic images of grain A7_60 are shown in Figure 11.
53
Figure 11. Images of DOM 08006 area (from Area 7, Figure 1) containing C- and N-anomalous
grain A7_60 (white boxes). a) NanoSIMS 12C image, b) secondary electron image, c)
NanoSIMS δ13C image, and d) NanoSIMS δ15N image. Delta-values are in ‰.
54
Figure 12. NanoSIMS images of DOM 08006 (left) and QUE 99177 (right). (a) and (b) 12C ion
images; units are counts per second. (c) and (d) In situ δD images; units in ‰. (e) and (f)
Purified IOM δD images; units in ‰. Note the similar intensity scales for 12C images but much
wider scales for δD in the QUE 99177 data compared to the DOM 08006 data. The apparent
isotopic variations in (c) and (e) are not statistically significant. QUE 99177 in situ data were
acquired previously with the Carnegie NanoSIMS under similar conditions to those reported
here for DOM 08006. Scale bars are 2 µm.
55
Figure 13. a) NanoSIMS 17O/16O map containing presolar grains DOM-17 and DOM-18. b) Bright-
field TEM image of FIB section containing presolar grains.
56
Figure 14. a) High Angle Annular Dark Field (HAADF) STEM image, b) STEM-based EDS
element maps, and c) selected-area electron diffraction pattern of grain DOM-18. The outline
of the grain is indicated on TEM image and EDS maps. The "RGB on HAADF" panel overlays
a false-color combination of the Mg, Si, and Fe maps (red, green, blue, respectively) on the
HAADF STEM image; the yellow color indicates grain is poor in Fe. The diffraction pattern
indicates that it is crystalline olivine, with specific lattice planes indicated for two diffraction
spots.
57
Figure 15. Transmission electron microscopy of presolar grain DOM-77: a) Dark field STEM
image. b) RGB STEM-EDS overlay on the STEM image. c) HR-TEM and Fast Fourier
Transform (FFT) of Mg-rich olivine region showing lattice spacings of 0.251 nm, 0.225 nm,
and 0.205 nm. d) HR-TEM and Ca-rich region with lattice spacings 0.406 nm and 0.367 nm;
FFT inset is from the area indicated by the box. Grain outline is indicated on a) and b).
58
Figure 16. Results of a Monte Carlo calculation to estimate the uncertainty in the derived
abundance of O-rich presolar grains in DOM 08006 (see text for details).
59
Figure 17. Abundances of presolar O-rich grains in primitive extraterrestrial materials plotted
against total area mapped by NanoSIMS. DOM (this work): DOM 08006 data reported here;
Acfer 094 (Vollmer et al., 2009b); Adelaide (Floss and Stadermann, 2012); A77: ALH 77307
(Nguyen et al., 2010); AMMs: Antarctic Micrometeorites (Yada et al., 2008); DOM (H18):
DOM 08006 matrix (Haenecour et al., 2018); E61: Elephant Moraine 92161 (Leitner et al.,
2016); DOM (comb): combined DOM 08006 dataset from the present study and that of
Haenecour et al. (2018); GRV: Grove Mountains 02170 (Zhao et al., 2013); G29: Graves
Nunataks 95229 (Leitner et al., 2016); IDPs: interplanetary dust particles (Floss et al., 2006;
Busemann et al., 2009); LAP: LAP 031117 matrix (Haenecour et al., 2018); MIL25: Miller
Range 07525 (Leitner et al., 2016); M00: MET 00426 (Floss and Stadermann, 2009a); NWA:
Northwest Africa 852 (Leitner et al., 2012b); Q99: QUE 99177 (Floss and Stadermann, 2009a;
Nguyen et al., 2010). Error bars are 1σ.
60
|
1706.06698 | 1 | 1706 | 2017-06-20T23:11:08 | Science enabled by a Moon Village | [
"astro-ph.EP",
"astro-ph.IM"
] | A human-robotic "Moon Village" would offer significant scientific opportunities by providing an infrastructure on the lunar surface. An analogy would be the way in which human outposts in Antarctica facilitate research activities across multiple scientific disciplines on that continent. Scientific fields expected to benefit from a Moon Village will include: planetary science, astronomy, astrobiology, life sciences, and fundamental physics. In addition, a Moon Village will help develop the use of lunar resources, which will yield additional longer-term scientific benefits. | astro-ph.EP | astro-ph | INTERNATIONAL ACADEMY OF ASTRONAUTICS
10th IAA SYMPOSIUM ON THE FUTURE OF SPACE EXPLORATION:
TOWARDS THE MOON VILLAGE AND BEYOND
Torino, Italy, June 27-29, 2017
SCIENCE ENABLED BY A MOON VILLAGE
Ian A. Crawford
Department of Earth and Planetary Sciences, Birkbeck College, University of London, UK
[email protected]
Abstract: A human-robotic 'Moon Village' would offer significant scientific
opportunities by providing an infrastructure on the lunar surface. An analogy would be
the way in which human outposts in Antarctica facilitate research activities across
multiple scientific disciplines on that continent. Scientific fields expected to benefit
from a Moon Village will include: planetary science, astronomy, astrobiology, life
sciences, and fundamental physics. In addition, a Moon Village will help develop the
use of lunar resources, which will yield additional longer-term scientific benefits.
Keywords: Moon Village; Lunar science, Lunar exploration, Lunar resources
1. INTRODUCTION
The ESA Director General has suggested that the creation of a human-robotic lunar outpost (or 'Moon village')
would be a logical next step in human space exploration [1]. The creation of such an outpost would offer
significant scientific opportunities by providing an infrastructure on the lunar surface [2-5]. An analogy would be
the way in which human outposts in Antarctica facilitate research activities across multiple scientific disciplines
on that continent [6,7]. In the Moon Village case, scientific fields that might be expected to benefit will include:
planetary science, astronomy, astrobiology, life sciences, and fundamental physics. In addition, a Moon Village
would help assess the economic potential of lunar resources [8,9], while also acting as a 'market' for such
resources.
2. SCIENCE ENABLED BY A MOON VILLAGE
The extent to which different scientific areas will benefit from a lunar outpost will, in part, depend on its location
and on the duration for which it is occupied. Some of the scientific areas discussed below are more dependent on
these factors than others. For example, science questions related to the lunar poles would clearly benefit more
from a polar Moon Village than an equatorial one. Nevertheless, the kind of transportation and other infrastructure
required to establish a Moon Village is likely to facilitate human and robotic operations at locations that may not
be local to the Moon Village itself. Moreover, in the fullness of time we might envisage multiple such outposts
being established at different locations.
With these caveats in mind, we here address the major areas of science that would benefit from the scientific
infrastructure represented by a Moon Village.
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
2.1. Planetary science
As discussed in references [4,5], the lunar geological record still has much to tell us about the earliest history of
the Solar System, the origin and evolution of the Earth-Moon system, the geological evolution of rocky planets,
and the near-Earth cosmic environment throughout Solar System history. Accessing this record would be greatly
enabled by again having humans operating on the lunar surface, and especially if supported by a scientific
infrastructure such as that envisaged for the Moon Village. This would be especially true with regard to the range
of rock and soil samples that could be collected, analysed (and, if necessary, returned to Earth), and for the
implementation of complex exploratory activities such as sub-surface drilling.
Specific planetary science activities that would benefit from such an infrastructure include the following (see
references [4,5,10,11], and references cited therein, for additional details):
• Constraining the bombardment history of the inner solar system. The lunar cratering rate is used to
estimate the ages of cratered surfaces throughout the solar system, but it is not well calibrated (e.g. [12]).
Improving this calibration is a key objective of planetary science, and will require sampling impact melt
deposits from multiple (ideally hundreds) of impact craters of a wide range of ages. In addition,
recovering fragments of the impacting bodies will reveal how and if the population of impactors has
changed with time [13]. Accessing these materials will require a significant transportation infrastructure
(e.g. pressurised rovers) on the lunar surface, a significant sample return capacity and, ideally, in situ
analytical instruments.
• Understanding the impact cratering process. Impact cratering is a fundamental planetary process, but our
understanding of it is limited by lack of access to pristine craters of a wide range of sizes. The lunar
surface clearly exhibits the required diversity of craters. Using these craters to improve our knowledge of
impact cratering processes will require visiting many craters of a wide range of sizes and ages to conduct
sampling, geophysical surveys and, possibly, sub-surface drilling.
• Determining the structure, composition and evolution of the lunar interior. The Moon provides an
excellent example of a small rocky planet which largely preserves its internal structure from shortly after
its formation. As such it can provide insights into the early geological evolution of larger and more
complicated planets. To better understand the structure of the lunar interior will require emplacement of
geophysical tools (e.g. seismic networks and magnetic surveys), as well as sampling a wide range of lunar
crustal and volcanic rocks of diverse ages and compositions. Such studies will also aid in constraining
theories of lunar origin.
• Determining the extent and distribution of lunar volatiles. There is now convincing evidence that water
ice exists within permanently shadowed polar craters, and that hydrated materials also exist at high-
latitude (but not permanently shadowed) localities. In addition to their possible value as resources (see
below), such deposits will inform our knowledge of the role comets and meteorites have played in
delivering volatile substances to the terrestrial planets. Determining the nature and extent of these
volatiles will require surface activities (including sampling and near-surface drilling) at polar and high-
latitude locations.
• Understanding regolith processes on airless bodies. The lunar surface is a natural laboratory for
understanding space weathering and regolith processes throughout the solar system. Access to regoliths of
2
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
different composition, thickness, and latitude will be required to improve our knowledge of these
processes.
• Serendipitous discoveries. Humans are unique in our ability to recognize new observations or phenomena
to be of importance, even if not anticipated in advance. It follows that having humans living and working
on the lunar surface for long periods is likely to result in unanticipated discoveries that might not
otherwise be made. Although by definition unquantifiable, such discoveries may ultimately prove to be
among the most significant to result from a lunar outpost.
2.3. Astronomy
The lunar surface is a potentially useful platform for astronomical observations [14-16], and establishing the
equipment required for such observations would be facilitated by a human and robotic infrastructure on the Moon.
Key aspects include:
• Low frequency radio astronomy from the lunar farside. Radio wavelengths longer than about 10m cannot
penetrate the Earth's ionosphere, yet much valuable scientific information is expected to lie in this
frequency range (e.g. [16]). The lunar farside, which is permanently shielded from the Earth, is probably
the best location in the inner solar system for such observations and a human infrastructure on the Moon
would assist in the setting up of the necessary equipment (although care will also be required to ensure
that human operations do not degrade the natural radio-quietness of the location).
• Optical and infra-red astronomy. Although the lunar surface could in principle provide a platform for
optical and infra-red telescopes, there is a consensus that free-flying satellites (e.g. at the Earth-Sun L2
point) probably provide better platforms for such activities [17]. It is certainly true that some aspects of
the lunar surface environment (e.g. dust and, in most locations, extreme diurnal temperature ranges) are
not amenable to astronomical observations at these wavelengths. Nevertheless, access to a scientific
infrastructure, able to emplace, maintain and upgrade such instruments (the value of which has been
demonstrated by multiple HST servicing missions) might compensate for these disadvantages. Moreover,
the stability of the lunar surface might be enabling in the context of building optical/IR interferometric
arrays, and permanently shadowed polar craters might enable passive cooling of IR instruments. The pros
and cons of the lunar surface for optical/IR astronomy still need to be properly assessed and a lunar
outpost would at least facilitate such an assessment.
• Cosmic-ray observations. As the lunar surface is not shielded by either an atmosphere or a magnetic field
(other than the Sun's heliospheric magnetic field) it is an attractive location from which to study the flux
and composition of primary cosmic rays. This can also be done in free space, but the existence of a
human and robotic infrastructure on the lunar surface may facilitate the installation of the required
instrumentation. A similar argument holds for Gamma- and X-Ray observations.
2.3. Astrobiology
Astrobiology is usually defined as the study of the origin, evolution, distribution, and future of life in the universe.
As the Moon has (almost certainly) never had any indigenous life of its own, at first sight a scientific lunar
infrastructure might not seem especially relevant to astrobiology. However, we can identify the following areas of
astrobiological research that would benefit from lunar surface operations (see references [4,18,19], and references
cited therein, for additional details):
3
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
• Constraining the impact regime under which life arose and evolved on Earth. Understanding the extent to
which large meteorite impacts may have affected the origin and early evolution of life on Earth is a key
aspect of astrobiology. It will be naturally addressed as part of the wider calibration of inner solar system
cratering rate discussed above, and will require sampling of ancient impact melt deposits in the age range
4.5-3.8 Gyr ago.
• Accessing the records of solar and galactic evolution recorded in buried palaeoregolith layers. As
discussed in [18,20,21], buried ancient regoliths ('palaeoregoliths') are expected to contain records of the
solar wind and galactic cosmic ray (GCR) fluxes throughout solar system history, both of which will
inform our understanding of the past habitability of the Earth. Accessing such deposits will require
extensive fieldwork and, possibly, drilling to depths of tens of metres, and is the kind of large-scale
exploratory activity that would be greatly facilitated by a human infrastructure on the Moon.
• Sampling materials from early Earth, Mars and Venus. As pointed out in references [22,23], the lunar
surface may contain fragments of the early Earth (and possibly also of Mars and Venus) that pre-date the
oldest existing surfaces on those planets. Finding such material, especially for the early Earth, could help
constrain the timing of the origin and early evolution of life on our planet and would be of considerable
astrobiological significance. Finding such materials on the Moon would probably require extensive
exploratory activities [24] and would be aided by a human infrastructure such as would be provided by a
Moon Village.
• Sampling of Earth's early atmosphere. There is a possibility that molecules derived from the Earth's
atmosphere may be preserved in the lunar regolith [25,26]. If these can be identified in, and extracted
from, ancient palaeoregolith layers then they have the potential to provide insights into the evolution of
Earth's atmosphere with time. Accessing them will require the same sort of capabilities required for
extracting solar wind and GCR records from palaeregolith layers as discussed above.
• Understanding the astrobiological significance of irradiated ices. Lunar polar ices exposed to GCRs may
be expected to form organic molecules [27]. Sampling of polar ices, as described above, will indicate the
extent to which this process occurs and its relevance to abiotic syntheses of organic molecules in
planetary and interstellar ices.
• Studying the survival of organic material (including spores and, conceivably, even live micro-organisms)
within the remains of space vehicles (including Apollo lunar module ascent stages) that have crashed onto
the lunar surface [28]. Sampling such localities, and determining what, if any, viable organic material has
survived for decades on the lunar surface will provide key information regarding the survival of life in the
space environment relevant to fundamental biology, planetary protection and panspermia.
• SETI from the Moon. At the other extreme of the astrobiology spectrum, a scientific infrastructure on the
Moon could help constrain the prevalence of technological civilisations in the Galaxy. There are two main
possibilities; (i) SETI could make us of radio astronomical facilities established on the Moon [29]; and,
more speculatively (ii) the lunar surface might be searched for artefacts of extraterrestrial origin [30,31].
Both would benefit from a human-tended infrastructure on the lunar surface, but the latter (speculative
though it is) would probably absolutely require such an infrastructure owing the large surface areas that
would need to be searched to place any meaningful limits on the existence of any such artefacts.
4
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
• Preparing for deep-space exploration. Insofar as astrobiology, broadly considered, includes consideration
of the future of life in the universe the prospects for human expansion beyond Earth fall within its remit.
Clearly establishing a Moon Village would in itself be a significant step in this direction. Moreover, it
would help build knowledge and expertise that will be required if humans are to move further out into the
solar system, and especially to the planet Mars. Much of this knowledge and experience will be in the
field of the life sciences, to which we now turn.
2.4. Life sciences
The lunar surface will provide multiple opportunities for research in the life sciences (e.g. [4,32,33]). Most of
these would not be dependent on the physical location of a human-tended outpost, although some would require a
prolonged presence. Gronstal et al. [32] have reviewed the kind of laboratory equipment that will be required for
such investigations, and it is clear that a human-tended outpost would greatly facilitate this work (and will, of
course, be essential for human studies). Specific areas of lunar life science investigations include:
• Understanding the response of life to the low, but non-zero, gravity. Although a lot of research has now
been performed on the response of living things to microgravity (e.g. [34,35]), no comparable data exists
for prolonged exposure to low, but non-zero, gravity. A lunar outpost would permit such studies on life-
forms ranging from individual cells to entire organisms (including humans). Fundamental insights into
biological processes may be expected from such studies, in addition to knowledge that will aid in the
future exploration of space.
• Understanding the response of life (again ranging from individual cells to entire organisms) to the lunar
radiation and dust environment.
• Understanding the response of human physiology to the lunar environment. Specifically human-oriented
research will aid in the development of countermeasures to help enable the long-term human habitation of
the Moon and other planetary surfaces [35].
• Agricultural experiments. If humans are to have a long-term future on the Moon, and on other planetary
surfaces, then growing food will at some stage become necessary. A lunar outpost would facilitate the
necessary research.
2.5. Fundamental physics
Although not a major driver for lunar exploration, several areas of fundamental physics research may benefit from
the scientific infrastructure represented by a lunar outpost. These include:
• Test of General Relativity through improved lunar laser ranging measurements [36].
• Tests of quantum entanglement over the Earth-Moon baseline [37].
• Searches for exotic particles (e.g. strange quark matter, dark matter, etc) interacting with the lunar
surface [38].
3. LUNAR RESOURCES
5
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
There is growing interest in the future use of lunar resources, both to support lunar exploration itself and as a
contribution to a developing space economy [9]. A self-sustaining Moon Village will very likely rely on
lunar resources (e.g. locally derived water and oxygen [8,9]), but equally the infrastructure provided by an
outpost of this kind will facilitate the prospecting for additional resources. Indeed, a Moon Village could help
kick-start a space economy by providing a market for commercial space resource companies while at the
same time providing infrastructure to support their activities. Although not strictly a scientific benefit of a
Moon Village per se, there is little doubt that, in the longer term, science will benefit from the development
of a space economy built on the use of space resources [39-41].
4. CONCLUSION
By analogy with scientific outposts in Antarctica, a human-robotic outpost on the Moon would enable multiple
scientific opportunities owing to the lunar surface infrastructure that it would provide. Scientific fields that would
be expected to benefit from such an infrastructure include: planetary science, astronomy, astrobiology, the life
sciences, and fundamental physics. In addition, a Moon Village would help initiate the utilization of lunar
resources, and perhaps help kick-start a space economy. Although beyond the scope of this paper, a Moon Village
would also offer important societal and cultural benefits to humanity, especially if, as envisaged, it is developed as
a truly global endeavor [42,43]. Moreover, although not explicitly identified as an objective in the current ISECG
Global Exploration Roadmap [44], an international Moon Village would clearly be consistent with this roadmap,
and more especially with the over-arching aims of the Global Exploration Strategy [45]. The scientific
community, and indeed the world community more generally, has the greatest possible interest in seeing this
initiative succeed.
REFERENCES
[1] Woerner, J., Foing, B., 'The "Moon Village" concept and initiative', Annual Meeting of the Lunar Exploration Analysis
Group, held 1-3 November, 2016 in Columbia, Maryland. LPI Contribution No. 1960, id.5084, (2016).
https://www.hou.usra.edu/meetings/leag2016/pdf/5084.pdf
[2] Taylor, G.J. 'The need for a lunar base: answering basic questions about planetary science'. In: Lunar Bases and Space
Activities of the 21st Century (ed. W.W. Mendell), Houston, TX; 189–197, (1985).
[3] Spudis, P.D, The Once and Future Moon. Washington, DC: Smithsonian Institution Press, (1996).
[4] Crawford, I.A., et al., 'Back to the Moon: the scientific rationale for resuming lunar surface exploration', Planet. Space
Sci., 74, 3–14, (2012).
[5] Crawford, I.A., Joy, K.H., 'Lunar exploration: opening a window into the history and evolution of the inner solar system',
Phil. Trans. Royal Soc., A372: 20130315, 1-21, (2014).
[6] Ehrenfreund, P. et al., 'Toward a global space exploration program: a stepping stone approach'. Adv. Space Res., 49, 2–
48, (2012).
[7] McKay, C., 'The case for a NASA research base on the Moon'. New Space, 1, 162-166, (2013).
[8] Anand, M., et al. 'A brief review of chemical and mineralogical resources on the Moon and likely initial in situ resource
utilisation (ISRU) applications', Planet. Space Sci., 74, 42–48, (2012).
[9] Crawford, I.A., 'Lunar resources: a review', Progress in Physical Geography, 39, 137-167, (2015).
[10] NRC, The Scientific Context for Exploration of the Moon. National Research Council, National Academies Press,
Washington DC, (2007).
[11] Neal, C.R., et al., 'Developing the global exploration roadmap: An example using the humans to the lunar surface
theme', Space Policy, 30, 156-162, (2014).
[12] Robbins, S.J., 'New crater calibrations for the lunar crater-age chronology', Earth Planet. Sci. Lett., 403, 188-198,
(2014).
[13] Joy, K.H., et al., 'The Moon: an archive of small body migration in the solar system', Earth Moon Planets, 118, 133-158
(2016).
[14] Burns, J.O., et al., 'Observatories on the Moon', Scientific American, 262(3), 18-25, (1990)
[15] Crawford, I.A., Zarnecki, J., 'Astronomy from the Moon', Astronomy and Geophysics, 49, 2.17-2.19, (2008).
[16] Jester, S., Falcke, H., 'Science with a lunar low-frequency array: from the dark ages of the Universe to nearby
exoplanets', New Astron. Rev., 53, 1-26, (2009).
6
10th IAA Symposium on the future of space exploration: towards the Moon Village and beyond
[17] Lester, D.F., et al., 'Does the lunar surface still offer value as a site for astronomical observatories?', Space Policy, 20,
99-107, (2004).
[18] Crawford, I.A., 'The astrobiological case for renewed robotic and human exploration of the Moon', Internat. J.
Astrobiology, 5, 191-197, (2006).
[19] Crawford, I.A., Cockell, C.S., 'Astrobiology on the Moon', Astronomy & Geophysics, 51, 4.11-4.14, (2010).
[20] Spudis, P.D., The Once and Future Moon, Smithsonian Institution Press, Washington D.C, (1996).
[21] Crawford, I.A., et al., 'Lunar palaeoregolith deposits as recorders of the galactic environment of the Solar System and
implications for astrobiology', Earth Moon Planets, 107, 75-85, (2010).
[22] Armstrong, J.C., et al., 'Rummaging through Earth's attic for remains of ancient life', Icarus, 160:183-196, (2002).
[23] Gutiérrez, J.L., 'Terrene meteorites in the moon: its relevance for the study of the origin of life in the Earth', ESA SP-
518, 187 – 191, (2002).
[24] Crawford, I.A., et al., 'On the survivability and detectability of terrestrial meteorites on the Moon', Astrobiology, 8,
242-252, (2008).
[25] Ozima, M., et al., 'Toward prescription for approach from terrestrial noble gas and light element records in lunar soils
understanding early Earth evolution', Proc. Nat. Acad. Sci. 105,17654–17658, (2008).
[26] Terada, K., et al., 'Biogenic oxygen from Earth transported to the Moon by a wind of magnetospheric ions', Nature
Astronomy, 1, 26, (2017).
[27] Lucey, P.G., 'Potential for prebiotic chemistry at the poles of the Moon', Proc. SPIE, 4137, 84-88, (2000).
[28] Glavin, D.P., et al., 'In situ biological contamination studies of the Moon: implications for planetary protection and life
detection missions', Earth Moon Planets, 107, 87-93, (2010).
[29] Maccone, C., 'SETI on the Moon'. In: Seckbach, J., et al. (eds.) Life in the Universe, Springer, pp. 303-306, (2004).
[30] Arkhipov, A.V., 'Earth-Moon system as a collector of alien artefacts', J. Brit. Interplanet. Soc., 51, 181-184, (1998).
[31] Rose, C., Wright, G., 'Inscribed matter as an energy-efficient means of communicating with an extraterrestrial
civilisation', Nature, 431, 47-49, (2004).
[32] Gronstal, A., et al., 'Lunar astrobiology: a review and suggested laboratory equipment', Astrobiology, 7, 767-782,
(2007).
[33] Cockell, C.S., 'Astrobiology – what can we do on the Moon?', Earth Moon Planets, 107, 3-10, (2010).
[34] Seibert, G., et al. (eds.), A World Without Gravity, ESA SP-1251, (2001).
[35] Horneck, G., et al. (eds.), HUMEX: Study on the Survivability and Adaptation of Humans to Long-Duration Exploratory
Missions, ESA SP-1264, (2003).
[36] Currie, D.G., et al., 'A lunar laser retroreflector for the 21st Century', AGU Abstract P51C-1467, (2010).
[37] Schneider, J., 'Test of quantum physics from Earth-Moon correlations', EPSC Abstract 202, (2010).
[38] Banerdt, W.B., et al., 'Using the Moon as a low-noise seismic detector for strange quark nuggets', Nuclear Physics B
Proceedings Supplements, 166, 203-208, (2007).
[39] Crawford, I.A., 'The long-term scientific benefits of a space economy', Space Policy, 37, 58-61, (2016).
[40] Elvis, M., 'What can space resources do for astronomy and planetary science?', Space Policy, 37, 65-76, (2016).
[41] Metzger, P.T., 'Space development and space science together, an historic opportunity', Space Policy, 37, 77-91, (2016).
[42] Crawford, I.A., 'Space development: social and political implications', Space Policy, 11, 219-225, (1995).
[43] White, F., The Overview Effect: Space Exploration and Human Evolution, 3rd edition, AIAA, (2014).
[44] International Space Exploration Coordination Group (ISECG), The Global Exploration Roadmap. (2013)
http://www.globalspaceexploration.org/web/isecg/news/2013-08-20.
[45] The Global Exploration Strategy: Framework for Coordination (2007).
http://esamultimedia.esa.int/docs/GES_Framework_final.pdf.
7
|
1301.3005 | 1 | 1301 | 2013-01-14T15:08:42 | Physical properties of the WASP-44 planetary system from simultaneous multi-colour photometry | [
"astro-ph.EP"
] | We present ground-based broad-band photometry of two transits in the WASP-44 planetary system obtained simultaneously through four optical (Sloan g', r', i', z') and three near-infrared (NIR; J, H, K) filters. We achieved low scatters of 1-2 mmag per observation in the optical bands with a cadence of 48 s, but the NIR-band light curves present much greater scatter. We also observed another transit of WASP-44 b by using a Gunn-r filter and telescope defocussing, with a scatter of 0.37 mmag per point and an observing cadence around 135 s. We used these data to improve measurements of the time of mid-transit and the physical properties of the system. In particular, we improved the radius measurements of the star and planet by factors of 3 and 4, respectively. We find that the radius of WASP-44 b is 1.002 R_Jup, which is slightly smaller than previously thought and differs from that expected for a core-free planet. In addition, with the help of a synthetic spectrum, we investigated the theoretically-predicted variation of the planetary radius as a function of wavelength, covering the range 370-2440 nm. We can rule out extreme variations at optical wavelengths, but unfortunately our data are not precise enough (especially in the NIR bands) to differentiate between the theoretical spectrum and a radius which does not change with wavelength. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 13 (2012)
Printed 13 August 2018
(MN LATEX style file v2.2)
Physical properties of the WASP-44 planetary system from
simultaneous multi-colour photometry
L. Mancini,1⋆ N. Nikolov,1 J. Southworth,2 G. Chen,1,3 J. J. Fortney,4
J. Tregloan-Reed,2 S. Ciceri,1 R. van Boekel,1 and Th. Henning1
1Max Planck Institute for Astronomy, Konigstuhl 17, 69117 -- Heidelberg, Germany
2Astrophysics Group, Keele University, Newcastle-under-Lyme, ST5 5BG, UK
3Purple Mountain Observatory & Key Laboratory for Radio Astronomy, 2 West Beijing Road, Nanjing 210008, China
4Department of Astronomy & Astrophysics, University of California, Santa Cruz, CA 95064, USA
Accepted. Received.
ABSTRACT
We present ground-based broad-band photometry of two transits in the WASP-44
planetary system obtained simultaneously through four optical (Sloan g ′, r′, i′, z ′)
and three near-infrared (NIR; J, H, K) filters. We achieved low scatters of 1 -- 2 mmag
per observation in the optical bands with a cadence of ≈ 48 s, but the NIR-band light
curves present much greater scatter. We also observed another transit of WASP-44
b by using a Gunn-r filter and telescope defocussing, with a scatter of 0.37 mmag
per point and an observing cadence around 135 s. We used these data to improve
measurements of the time of mid-transit and the physical properties of the system. In
particular, we improved the radius measurements of the star and planet by factors of 3
and 4, respectively. We find that the radius of WASP-44 b is 1.002 ± 0.033 ± 0.018 RJup
(statistical and systematic errors, respectively), which is slightly smaller than previ-
ously thought and differs from that expected for a core-free planet. In addition, with
the help of a synthetic spectrum, we investigated the theoretically-predicted variation
of the planetary radius as a function of wavelength, covering the range 370 -- 2440 nm.
We can rule out extreme variations at optical wavelengths, but unfortunately our
data are not precise enough (especially in the NIR bands) to differentiate between the
theoretical spectrum and a radius which does not change with wavelength.
Key words: stars: planetary systems -- stars: fundamental parameters -- stars: indi-
vidual: WASP-44
3
1
0
2
n
a
J
4
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
0
0
3
.
1
0
3
1
:
v
i
X
r
a
1
INTRODUCTION
The transit method is not only an excellent technique to
detect extrasolar planets, but also provide unrivalled access
to their physical parameters. The geometry of a transiting
extrasolar planet (TEP) system permits measurement of
the mass, radius and density of a planet. These quantities
require some input from stellar evolutionary theory, but the
planetary surface gravity can be measured directly. A TEP
system is therefore a potential horn of plenty of physical
information on planetary systems (Seager & Sasselov 2000;
Sudarsky et al. 2000; Brown et al. 2001; Hubbard et al.
2001;
Sudarsky et al.
2003; Charbonneau et al. 2005; Holman & Murray 2005;
Winn et al. 2005; Southworth et al. 2007).
Seager & Mall´en-Ornelas
2003;
Currently, most of the TEPs discovered by ground-
based transit surveys (e.g. Hellier et al. 2012; Smalley et al.
⋆ [email protected]
2012; Bakos et al. 2012; Hartman et al. 2012; Penev et al.
2013; Bryan et al. 2012; Beatty et al. 2012; Siverd et al.
2012) are close-in hot Jupiters, because the detection prob-
ability for such planets is much greater than for those
which are smaller or on wider orbits. Hot Jupiters are also
relatively straightforward to detect via the radial velocity
method, as they induce comparatively large velocity vari-
ations in their host stars. The existence of hot Jupiters
was unexpected (Mayor and Queloz 1995), and led to the
birth of new formation and migration theories (see the re-
views of D'Angelo et al. 2010 and Lubow and Ida 2010).
The bias of the ground-based surveys in favour of this class
of planets is clear, whereas the Kepler satellite has already
found hundreds of smaller and longer-period planet candi-
dates from only 16 months of data (Borucki et al. 2011a,b;
Batalha et al. 2012).
Another fascinating opportunity offered by TEPs is the
possibility to probe their atmospheres. Due to atomic and
molecular absorption, the analysis of the spectrum of the
2
L. Mancini et al.
parent stars during transit (primary eclipse) is an excellent
way to probe their atmospheric composition. The existence
of two different classes of hot Jupiters (pM and pL) has been
suggested according to the incident stellar flux received from
the parent star and to the expected amount of absorbing
substances, such as gaseous titanium oxide (TiO) and vana-
dium oxide (VO), in their atmospheres (Fortney et al. 2008).
According to theoretical models, these two oxidized elements
should be extremely strong absorbers for the pM class be-
tween 450 nm and 700 nm, causing such planets to have a hot
(∼2000 K) stratosphere. An observable variation of radius
with wavelength, due to the opacities of these elements, is
consequently expected in the optical bands (Burrows et al.
2007). The radius of a pM planet should be 3% lower at
350 -- 400 nm versus 500 -- 700 nm. The radius variations for the
colder pL-class planets (incident flux <109 erg s−1 cm−2) are
smaller, but a significant contribution to the opacity in the
optical band is expected from Na I at ∼590 nm, and K I at
∼770 nm (Fortney et al. 2008, 2010; Burrows et al. 2010).
Transmission spectroscopy is essential for detecting atmo-
spheric absorbers, and is complementary to observations at
secondary eclipse which can help to determine the day-side
atmospheric temperature structure.
Accurate measurements of planet radii provide critical
constraints for astrophysicists working on planet formation
and evolution, because they provide an indication of the
planet internal heat, structure and composition, as well as
the heating mechanism for those which undergo strong tidal
effects and stellar irradiation, allowing the discrimination
between competing theories (Pollack et al. 1996; Boss 1997).
In particular, the most irradiated hot Jupiters are charac-
terised by anomalously inflated radii for which the expla-
nation remains unclear (Bodenheimer et al. 2003; Gu et al.
2004; Gaudi 2005; Fortney et al. 2007; Jackson et al.
2008; Dobbs-Dixon & Lin
2009;
Ibgui et al.
Ibgui & Burrows
2010;
2011;
Demory & Seager 2011).
2011; Miller & Fortney
2008; Batygin et al.
2009; Miller et al.
2009;
Laughlin et al.
Photometric and spectroscopic analyses of planet at-
mospheres started with the use of optical, near-infrared
(NIR) and infrared (IR) instruments aboard the Hub-
ble and Spitzer space telescopes (Charbonneau et al. 2002;
Vidal-Madjar et al. 2003, 2004; Richardson et al 2006;
Tinetti et al. 2007; Knutson et al. 2007a,b; Richardson et al
2007; Swain et al. 2008; Pont et al. 2008; Beaulieu et al.
2008, 2010, 2011; Gillon et al. 2012) and led to the detection
of Rayleigh scattering in a few hot Jupiter atmospheres, plus
absorption lines associated with H2O, Na, CH4, TiO and VO
(see above references but also Ballester et al. 2007; Barman
2007; Sing et al. 2008a,b; Lecavelier des Etangs et al. 2008;
D´esert et al. 2008; Zahnle et al. 2009; Spiegel et al. 2009;
Fossati et al. 2010; Tinetti et al. 2010; D´esert et al. 2011;
Wood et al. 2011; Gibson et al. 2011; Crossfield et al. 2012;
Crouzet et al. 2012).
After a number of pioneering experiments with null
results, ground-based spectroscopy also obtained some in-
teresting results. Observations performed at the Hobby-
Eberly and Subaru telescopes, in the spectral range 500 --
900 nm, detected Na absorption in the transmission spec-
trum of HD 189733 b (Redfield et al. 2008) and HD 209458 b
(Snellen et al. 2008). NIR spectroscopy of HD 209458 b at
the VLT in the range 2291 -- 2349 nm reported a significant
wavelength shift in absorption lines from carbon monoxide
in the planet's atmosphere (Snellen et al. 2010).
High-resolution NIR spectroscopic measurements at the
Keck telescopes over the wavelength range 2100 -- 2400 nm,
revealed an upper limit for CO absorption in the atmosphere
of HD 209458 b (Deming et al. 2005), and that the super-
Earth GJ 1214 b is a H-dominated planet (Crossfield et al.
2011). GJ 1214 b was also studied through multi-object spec-
troscopy from 0.61 to 0.85 µm, and in the J, H, and
K atmospheric windows by using VLT/FORS and Magel-
lan/MMIRS, the data being consistent with a featureless
transmission spectrum for the planet (Bean et al. 2011).
A radius variation with wavelength was investigated
for HD 209458 b by Knutson et al. (2007a) using ten-colour
HST photometry. Different analyses of these data have given
conflicting results (Knutson et al. 2007a; Barman 2007;
Sing et al. 2008a; Southworth 2008). An HST transmis-
sion spectrum of HD 189733 b covering 270 -- 570 nm showed
a gradual increase of radius towards shorter wavelengths
which was interpreted as a result of Rayleigh scattering from
a high-altitude atmospheric haze (Sing et al. 2011a). But ob-
servations covering 550 -- 1050 nm did not provide any indica-
tion of the expected Na I or K I features (Pont et al. 2008),
and those in the ranges 1082 -- 1168 nm and 1514 -- 1693 nm did
not detect any variation in planetary radius (Gibson et al.
2012a).
Using the Gran Telescopio Canarias (GTC), Sing et al.
(2011b) claimed the first evidence for potassium in an extra-
solar planet, from photometry of the XO-2 system in four
narrow red-optical passbands, by detecting K I absorption
at 766.5 nm. This technique was also used by Col´on et al.
(2012) to study HD 80606 b, who found an unexplained large
(≈ 4.2%) change in the apparent planetary radius between
the wavelengths 769.9 and 777.4 nm. The differential spec-
trophotometry technique was used to obtain two transit light
curves of XO-2 b with GTC by Sing et al. (2012), who de-
tected significant absorption in the planetary atmosphere
in a 50-A bandpass centred on the Na I doublet. Instead,
the presence of an Na-rich atmosphere in WASP-29 b was
recently ruled out by Gibson et al. (2012b), using Gemini-
South GMOS transit spectrophotometry.
Southworth et al. (2012b) recently presented a study
of possible radius variations of the hot Jupiter HAT-P-5 b,
based on photometry obtained simultaneous in the u, g, r
and I passbands. The authors detected a gradual increase of
the radius between 450 nm and 850 nm, plus a substantially
larger planetary radius at 350 nm. The latter phenomenon
can be explained by systematic errors in the u-band pho-
tometry, but is also consistent with Rayleigh scattering as
in the case of HD 189733 b. A similar multi-band investiga-
tion of HAT-P-8 (Mancini et al. 2013) suggests the presence
of strong optical absorbers near the terminator of this TEP.
Here we focus our attention on the planetary system
WASP-44, discovered by Anderson et al. (2012). The system
consists of a 0.89 MJup planet on a 2.42-day orbit around an
inactive G8 V star (V = 12.9, [Fe/H] = +0.06). In this work
we present the first photometric follow-up since its discovery
was announced, covering seven optical/NIR passbands. We
refine the physical properties of the system and attempt to
probe for radius variations in these passbands.
Our paper is structured as follows. In § 2 we describe
the instruments used for the observations and give some
details concerning the data reduction. In § 3 we illustrate
the analysis of the data which led to the refinement of the
orbital period and the physical properties of the WASP-44
system. Then, we examine the variation of the planetary
radius with wavelength. Finally, in § 4 we summarise our
results.
2 OBSERVATIONS AND DATA REDUCTION
Two transits of WASP-44 b were recorded on 2011 October
2 and 6, using the Gamma Ray Burst Optical and Near-
Infrared Detector (GROND) instrument mounted on the
MPG1/ESO 2.2 m telescope at ESO La Silla, Chile. GROND
is an imaging system capable of simultaneous photometric
observations in four optical (identical to Sloan g′, r′, i′, z′)
and three NIR (J, H, K) passbands (Greiner et al. 2008).
Each of the four optical channels is equipped with a back-
illuminated 2048 × 2048 E2V CCD, with a field of view
(FOV) of 5.4′ × 5.4′ at a scale of 0.158′′/pixel. The three
NIR channels use 1024 × 1024 Rockwell HAWAII-1 arrays
with a FOV of 10′ × 10′ at 0.6′′/pixel.
We applied a slight telescope defocus and obtained
repeated integrations during both runs. Autoguiding was
used to keep the stars on the same pixels. All images were
digitised using the fast read-out mode (∼10 s) to improve
the time sampling. On 2011/10/02 we monitored the flux
of WASP-44 in nearly photometric conditions, and saw a
monotonical seeing decline from 0.50′′ to 1.03′′ as airmass
decreased from 1.08 to 1.05 and then rose to 1.54. Expo-
sures of 20 s were used in the optical channels and stacks
of four 4 s images were obtained in the NIR channels. On
2011/10/06 we experienced worse atmospheric conditions,
and the airmass changed from 2.07 to 1.05 during our obser-
vations. Images with exposure times of 25 s were obtained in
the optical channels, and stacks of six 4 s images in the NIR
channels. For the optical frames we selected a fast read-out
mode, which provided a 15 s readout. Due to the necessary
synchronization of the optical and NIR starting times for
each exposure, the observing cadence resulted to be of ≈ 48
and ≈ 25 s in the optical and NIR bands respectively.
full
Another
transit of WASP-44 b was observed
through a Gunn r filter on the night of 2011/09/02 using the
Wide Field Camera (WFC) on the 2.5 m Isaac Newton Tele-
scope (INT) at La Palma. WFC is an optical mosaic camera
consisting of four thinned EEV 2k×4k CCDs, with a plate
scale of 0.33′′ pixel−1. The telescope was heavily defocussed
so the point spread function (PSF) of the target and com-
parison stars had not more than 35 000 counts per pixel. We
used the fast readout mode to reduce readout time down to
about 15 s. An exposure time of 2 min per frame was used,
and the resulting observing cadence was roughly 135 s. The
night was photometric. We were not able to autoguide the
telescope as the autoguider is incorporated into the WFC
and so was also defocussed. A summary of the observational
data is reported in Table 1.
Reduction of the optical frames was undertaken using
standard methods. We created master bias and flat-field
images by median-combining sets of bias images and sky
1 Max Planck Gesellschaft.
Physical properties of WASP-44 b
3
Table 2. Excerpts of the light curves of WASP-44. The full
dataset will be made available at the CDS.
Telescope
Filter
BJD (TDB)
Diff. mag.
Error
INT
Gunn r
INT
Gunn r
ESO 2.2m Sloan g′
ESO 2.2m Sloan g′
ESO 2.2m Sloan r′
ESO 2.2m Sloan r′
ESO 2.2m Sloan i′
ESO 2.2m Sloan i′
ESO 2.2m Sloan z′
ESO 2.2m Sloan z′
2455807.576418
2455807.579370
2455836.647443
2455836.649992
2455836.647443
2455836.649992
2455836.647443
2455836.649992
2455836.647443
2455836.649992
0.00036
0.00009
0.00126
0.00210
-0.00097
0.00079
0.00096
0.00138
-0.00143
0.00100
0.00041
0.00042
0.00100
0.00123
0.00088
0.00109
0.00109
0.00134
0.00142
0.00179
flats, and used them to correct the science images. Aperture
photometry was performed using the idl2/astrolib3 im-
plementation of daophot (Stetson 1987; Southworth et al.
2009). The apertures were placed manually and shifted to
account for pointing variations, which were measured by
cross-correlating each image against a reference image. We
experimented with wide range of aperture sizes and retained
those which gave photometry with the lowest scatter com-
pared to a fitted model. The times of observation were con-
verted from UTC to BJD(TDB) using the IDL procedures
of Eastman et al. (2010).
Differential photometry was obtained in each filter using
between two and four comparison stars, two of which were
enough bright to produce count rates comparable to those
of WASP-44. All good comparison stars were combined into
one ensemble by weighted flux summation. Slow variations
in the apparent brightness of the reference stars, primarily
attributable to atmospheric effects, were treated by fitting
a polynomial to regions outside transit, whilst simultane-
ously optimising the weights of the comparison stars. Given
the limited number of comparison stars in the small field
of view of GROND, and the shape of the slow brightness
variations, we used a second-order polynomial versus time.
Uncertainties introduced by this procedure were considered
in the modelling of the light curves, as described in the next
section.
The NIR frames were also calibrated in a standard way,
including dark subtraction, flat correction and sky subtrac-
tion. The master sky images used in sky subtraction were
created from two sets of 20-position dithering sky measure-
ments, one before the science observation and one after. We
performed aperture photometry on the calibrated NIR im-
ages as well. The aperture locations were determined us-
ing idl/find. Various combinations of aperture and annulus
sizes were check to find the best photometry. We also care-
fully made ensembles of comparison stars, and chose the
group which showed the least deviation from the target. Af-
ter normalising the target light curve with the composite
reference light curve, we decorrelated the data with posi-
tion, seeing and airmass in order to remove the correlated
2 The acronym idl stands for Interactive Data Language and
is a trademark of ITT Visual Information Solutions. For further
details see http://www.ittvis.com/ProductServices/IDL.aspx.
3 http://idlastro.gsfc.nasa.gov/
4
L. Mancini et al.
Table 1. Log of the observations presented in this work. Nobs is the number of observations and "Moon illum." is the fractional illumination
of the Moon at the midpoint of the transit. The aperture sizes are the radii of the software apertures for the star, inner sky and outer sky,
respectively. β is the factor used to inflate the errorbars (see § 3). Transit #1 was observed using the 2.5 m INT in La Palma, whereas
transits #2 and #3 using the MPG/ESO 2.2 m telescope in La Silla.
Transit
Date
Start/end Time Nobs Exposure
time (s)
(UT)
Filter
Airmass
Moon
illum.
Aperture
sizes (px)
Scatter
(mmag)
β
1
2
2
2
2
2
2
2
3
3
3
3
3
3
3
2011 09 03
2011 10 02
2011 10 02
2011 10 02
2011 10 02
2011 10 02
2011 10 02
2011 10 02
2011 10 06
2011 10 06
2011 10 06
2011 10 06
2011 10 06
2011 10 06
2011 10 06
00:28 - 00:57
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
03:21 - 07:35
156
317
317
317
317
631
631
631
317
313
313
313
623
623
623
120
20
20
20
20
3
3
3
25
25
25
25
3
3
3
Gunn r
Sloan g′
Sloan r′
Sloan i′
Sloan z′
J
H
K
Sloan g′
Sloan r′
Sloan i′
Sloan z′
J
H
K
1.60 → 1.32 → 2.18
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
1.08 → 1.05 → 1.54
2.07 → 1.05
2.07 → 1.05
2.07 → 1.05
2.07 → 1.05
2.07 → 1.05
2.07 → 1.05
2.07 → 1.05
25, 35, 50
16, 35, 55
17, 40, 60
20, 30, 45
12, 25, 40
31%
40%
40%
40%
40%
40% 6.0, 15.0, 23.0
40% 2.5, 11.5, 21.0
40% 3.0, 11.0, 21.0
80%
80%
80%
80%
80% 9.0, 17.0, 25.0
80% 6.5, 13.5, 22.0
80%
3.5, 4.5, 18.0
19, 30, 50
22, 40, 60
20, 40, 60
16, 40, 60
0.37
1.23
1.08
1.34
1.79
7.66
8.11
5.98
1.89
1.70
1.79
2.52
5.23
6.96
7.65
1.15
1.07
1.00
1.00
1.10
1.09
1.08
1.15
1.06
1.32
1.35
1.26
1.38
1.34
1.04
red noise. We extracted the optimal NIR light curves for the
two nights according to the rms of O -- C residuals and con-
sistency of the transit depth between two nights. Details are
given in Appendix A. The resulting photometry is given in
Table 2.
Although we have performed all the possible calibra-
tions and corrections, the scatter of the NIR data is much
larger than that of the optical data, and the NIR light curves
are heavily dominated by red noise. This problem is related
to the adopted observing strategy that unfortunately did
not allow a good SNR to be obtained4.
Our analysis also included the dataset presented by
Anderson et al. (2012), which was obtained through a Gunn
r filter and using EulerCam mounted on the 1.2 m Euler-
Swiss telescope at ESO La Silla.
3 ANALYSIS
We have measured the physical properties of the WASP-44
planetary system following the methodology of the Homoge-
neous Studies project (Southworth 2008, 2009, 2010, 2011,
2012). We refer the reader to those works for a detailed de-
scription of the approach.
Since the absolute values of the observational errors
from our pipeline (which come ultimately from the aper
subroutine) were found to be underestimated, we adopt the
standard practice of rescaling them for each dataset to give
a reduced χ2 of χ2
ν = 1. Then, in order to account for time-
correlated errors (i.e. red noise, which can significantly affect
ground-based data) and derive more realistic uncertainties,
we inflated the errorbars further by multiplying the data
4 After these first observations, we changed our observing strat-
egy adopting the defocussing technique for GROND. Thanks to
this approach, we can now obtain scatter smaller than 2 mmag
per observation in the NIR bands and smaller than 1 mmag in
the optical ones, without compromising the sampling (Nikolov et
al., in prep.).
weights by a factor β > 1. The β approach (e.g. Pont et al.
2006; Winn et al. 2007) is a widely used way to assess
red noise (e.g. Gillon et al. 2006; Winn et al. 2008, 2009;
Gibson et al. 2008; Nikolov et al. 2012; Southworth et al.
2012a,b). The factor β is a measurement of how close the
data noise is to the Poisson approximation, and is found
by binning the data and evaluating the ratio between the
size of the residuals versus what would be expected if the
data followed Poisson statistics. We evaluated values of the
β factor for each individual transit and for groups of 10 dat-
apoints; they are reported in Table 1. For a more exhaustive
discussion about rescaling errorbars, see Andrae (2010).
3.1 Period determination
As a first step, we fitted each light curve individually us-
ing the jktebop code (see § 3.2) in order to find the tran-
sit times. Their uncertainties were estimated from Monte
Carlo simulations. To these we added two timings obtained
by Anderson et al. (2012) and four obtained by amateur as-
tronomers, which are available on the TRESCA5 website
(see Table 3). We excluded from this analysis our NIR data
since they are much noisier than the optical ones. All timings
were placed on BJD(TDB) time system. The resulting mea-
surements of transit midpoints were fitted with a straight
line to obtain a new orbital ephemeris:
T0 = BJD(TDB)2 455 434.37642(37) + 2.4238133(23) × E
(1)
where E is the number of orbital cycles after the refer-
ence epoch (the midpoint of the first transit observed by
Anderson et al. 2012) and quantities in brackets denote the
uncertainty in the final digit of the preceding number. The
fit has χ2
ν = 2.42, so the uncertainties were increased to
account for this. Despite this large χ2
ν, we do not note any
5 The TRansiting ExoplanetS and CAndidates (TRESCA) web-
site can be found at http://var2.astro.cz/EN /tresca/index.php
Table 3. Central transit times of WASP-44 and their residuals
versus the ephemeris derived in this work. TRESCA refers to the
"TRansiting ExoplanetS and CAndidates" website.
Central transit time
BJD(TDB) -- 2400000
Cycle Residual Reference
no.
(JD)
55434.37637 ± 0.00040
55453.76639 ± 0.00042
55807.64374 ± 0.00013
55814.91655 ± 0.00150
55829.45489 ± 0.00245
55829.46151 ± 0.00163
55836.72905 ± 0.00020
55836.72979 ± 0.00030
55836.72900 ± 0.00020
55836.72928 ± 0.00015
55841.57719 ± 0.00035
55841.57757 ± 0.00046
55841.57684 ± 0.00028
55841.57769 ± 0.00031
56127.58624 ± 0.00048
0 -0.00005 Anderson et al. (2012)
8 -0.00054 Anderson et al. (2012)
0.00007 This work (INT r)
154
157
0.00144 Evans P. (TRESCA)
163 -0.00310 Lomoz F. (TRESCA)
163
0.00315 Lomoz F. (TRESCA)
166 -0.00038 This work (GROND g′)
0.00036 This work (GROND r′)
166
166 -0.00043 This work (GROND i′)
166 -0.00015 This work (GROND z′)
0.00014 This work (GROND g′)
168
0.00052 This work (GROND r′)
168
168 -0.00021 This work (GROND i′)
0.00064 This work (GROND z′)
168
286 -0.00078 Sauer T. (TRESCA)
systematic deviation from the predicted transit times. A plot
of the residuals around the fit is shown in Fig. 1. Since the
number of known observed transits of this planet is still very
low, no conclusions can be drawn regarding the existence of
any third body in the system.
3.2 Light curve modelling
The light curves were modelled using the jktebop6 code.
The primary fitted parameters were the sum and ratio of the
fractional radii of the star and planet, rA+rb and k = rb/rA,
and the orbital inclination, i. The fractional radii of the com-
ponents are defined as rA = RA/a and rb = Rb/a, where a
is the orbital semimajor axis, and RA and Rb are the true
radii of the two objects. Additional parameters of the fit in-
cluded the light level outside transit and the midpoint of the
transit. Limb darkening (LD) was imposed using a quadratic
law and the corresponding coefficients fixed at values theo-
retically predicted using model atmospheres (Claret 2004).
Southworth (2008) performed three different treatments of
the LD coefficients: (i) both coefficients fixed to theoreti-
cal values; (ii) both coefficients fitted; (iii) the linear LD
coefficient fitted and the nonlinear one fixed to its theoret-
ically predicted value. While the second treatment is possi-
ble only when the data are of high quality, the use of the
third alternative has a negligible impact on the final results.
Uncertainties were calculated using both Monte Carlo sim-
ulations and a residual-permutation algorithm (Southworth
2008), and the larger of the two values was retained for each
output quantity. The orbital eccentricity was fixed to zero
(Anderson et al. 2012). All the datasets were solved individ-
ually. As in the previous section, due to their large scatter
(see Table 1), we did not use the NIR data for estimating
the physical properties of the planetary system.
Representative photometric parameters were obtained
for each dataset and are given in Table 4. The corresponding
6 jktebop is written in FORTRAN77 and the source code is
available at http://www.astro.keele.ac.uk/ jkt/
Physical properties of WASP-44 b
5
INT
1.00
aeaeae
ae ae
aeaeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeaeae
aeae
aeaeaeae
ae
aeaeae
ae
Euler
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeaeae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
aeae
aeaeae
aeae
aeaeae
aeae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
aeae
ae
aeae
ae
aeaeae ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
aeae
ae
ae
aeae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeaeae
ae
ae
ae
ae
ae
ae
aeaeaeae
aeaeaeaeaeaeae
ae
ae
aeae
aeaeae
aeae
aeaeae
aeae
ae ae
ae
aeaeaeaeae
aeaeaeaeae
aeae
ae
aeaeae
ae
ae
aeaeaeaeaeaeae
ae
x
u
l
F
d
e
s
i
l
a
m
N
r
0.99
0.98
0.97
0.96
0.95
0.94
INT
Euler
ae
ae
aeaeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeaeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
aeae
ae
ae
ae
ae
ae
ae
ae
0.00
ae
ae
0.04
-0.04
-0.02
0.02
Orbital Phase
Figure 4. As for Fig. 2 but for the r−filter INT and Euler
(Anderson et al. 2012) light curves of WASP-44.
best fits are shown in Fig. 2 and 3 for the GROND data and
in Fig. 4 for the INT and Euler data. The final photomet-
ric parameters are the weighted mean of the values for each
dataset. The agreement between light curves is excellent: the
χ2
ν value of the agreement of the individual parameters with
respect to the weighted mean is smaller than 0.75 for all
except k, where we found a modestly larger χ2
ν of 1.64. We
have found this situation to be common during our work on
the Homogeneous Studies papers. Table 4 also shows a com-
parison with the results from Anderson et al. (2012), which
are in good agreement with ours but have much larger er-
rorbars.
3.3 Physical properties of the WASP-44 system
Following the approach described in Southworth (2009), the
estimation of the physical properties of the WASP-44 sys-
tem was performed making use of standard formula (e.g.
Hilditch 2001). We used the photometric parameters mea-
sured in § 3.2, and the velocity amplitude, effective temper-
ature, and metallicity of the star (KA = 138.8 ± 9.0 km s−1,
Teff = 5410 ± 150 K, [Fe/H]= +0.06 ± 10) measured by
Anderson et al. (2012). We interpolated within tabulations
from theoretical stellar models to find the best agreement be-
tween the observed and model-predicted Teff , and the mea-
sured rA and calculated RA/a. This yielded the best-fitting
mass, radius, surface gravity and mean density of the star
(MA, RA, log gA and ρA) and of the planet (Mb, Rb, gb
and ρb). We also calculated the orbital semi-major axis (a),
planetary equilibrium temperature (Teq), Safranov number
(Θ), and the evolutionary age of the star.
The uncertainties in the input parameters were prop-
agated into the output physical properties using a pertur-
6
L. Mancini et al.
0.006
0.004
0.002
0.000
-0.002
-0.004
-0.006
s
y
a
d
s
l
a
u
d
i
s
e
R
ç
ae
ç
ç
ae
ae
ae
ae
ae
ae
ae
ae
ç
0
50
100
150
Cycle number
200
250
Figure 1. O-C diagram of the mid-transit times of WASP-44 b versus a linear ephemeris. The timings in black are from this work, and
in grey are from Anderson et al. (2012). The timings obtained by amateur astronomers are plotted using open circles. The uncertainties
of our points have been rescaled (see text).
0.47 Μm
1.00
0.62 Μm
0.77 Μm
0.93 Μm
1.24 Μm
1.66 Μm
2.20 Μm
0.95
0.90
0.85
g¢
r¢
i¢
z¢
J
H
K
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
éé
é
é
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
ééé
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
éé
é
é
é
éé
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
éé
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
0.80
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
ééé
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
0.47 Μm
0.62 Μm
0.77 Μm
0.93 Μm
1.24 Μm
1.66 Μm
2.20 Μm
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ã
ã
ã
ã
ã
ã
é
é
é
1.00
0.95
0.90
0.85
g¢
r¢
i¢
z¢
J
H
K
x
u
l
F
d
e
s
i
l
a
m
r
o
N
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
éé
é
é
é
éé
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
éé
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
0.80
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
éé
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
ééé
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
é
ééééé
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
éé
ééé
é
é
é
é
ééé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
éé
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
éé
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
éé
é
é
-0.04
-0.02
0.00
0.02
0.04
-0.04
-0.02
0.00
0.02
0.04
Orbital Phase
Orbital Phase
Figure 2. First set (2011/10/02) of GROND light curves of WASP-44 compared to the best jktebop fits using the quadratic LD law.
The residuals of the fits are plotted at the base of the figure, offset from zero.
bation analysis (Southworth et al. 2005). Systematic errors
were assessed by comparing results from five different sets
of theoretical models (see Southworth 2010). The sets of
physical properties found using each set of stellar models
is shown in Table 5 and the final physical properties of the
WASP-44 system are given in Table 6. The results obtained
by Anderson et al. (2012) are less precise but are in good
agreement with our own. In particular, we improved the
measurement precisions of the radii of both the planet and
the parent star, by factors of 4 and 3 respectively. The radius
of WASP-44 b that we found (1.002 ± 0.033 ± 0.018 RJup) is
smaller than that measured in the discovery paper and does
not match the predicted radii of coreless hot-Jupiter plan-
ets as estimated by Fortney et al. (2007). According to their
0.47 Μm
1.00
g¢
r¢
i¢
z¢
J
H
K
1.00
0.95
0.90
0.85
0.80
x
u
l
F
d
e
s
i
l
a
m
r
o
N
0.62 Μm
0.77 Μm
0.93 Μm
ã
ã
1.24 Μm
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ããã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ãã
ãã
ãã
ã
ã
ã
ã
ã
ããã
ã
ãã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ãã
ãã
ã
ã
ãã
ã
ããã
ãã
ãã
ã
ã
ã
ããã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
éé
é
é
é
éé
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
éé
é
é
é
é
é
éé
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
1.66 Μm
2.20 Μm
Physical properties of WASP-44 b
7
0.47 Μm
0.62 Μm
0.77 Μm
0.93 Μm
ã
1.24 Μm
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ãã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ããã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ãã
ã
ã
ããã
ã
ã
ã
ã
ã
ãã
ã
ã
ããã
ã
ã
ã
ã
ãã
ã
ã
ãã
ãã
ãã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ãã
ãã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ãã
ããã
ãã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ãã
ã
ã
ã
ãã
ã
ãã
ãã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ãã
ã
ã
ã
ã
ã
ã
ãã
ã
ãã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ããã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
ã
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
éé
é
ééé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
éé
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
éé
éé
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
éé
é
é
é
é
é
éé
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
ééé
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
éé
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
é
éé
é
é
é
é
é
é
é
1.66 Μm
2.20 Μm
g¢
r¢
i¢
z¢
J
H
K
0.95
0.90
0.85
0.80
-0.04
-0.02
0.00
0.02
0.04
-0.04
-0.02
0.00
0.02
0.04
Orbital Phase
Orbital Phase
Figure 3. Second set (2011/10/06) of GROND light curves of WASP-44 compared to the best jktebop fits using the quadratic LD law.
The residuals of the fits are plotted at the base of the figure, offset from zero.
Table 4. Parameters of the fits to the ten light curves of WASP-44. The final parameters are the weighted mean of the result for the
GROND, INT and Euler light curves.
Source
rA + rb
k
GROND g′−band #1
GROND g′−band #2
GROND r′−band #1
GROND r′−band #2
GROND i′−band #1
GROND i′−band #2
GROND z′−band #1
GROND z′−band #2
INT r−band
Euler r−band
Final results
0.1368 ± 0.0056
0.139 ± 0.010
0.1268 ± 0.0045
0.1284 ± 0.0080
0.1363 ± 0.0051
0.1303 ± 0.0087
0.1306 ± 0.0075
0.123 ± 0.010
0.1285 ± 0.0029
0.141 ± 0.011
0.1228 ± 0.0023
0.1188 ± 0.0028
0.1175 ± 0.0014
0.1190 ± 0.0029
0.1205 ± 0.0016
0.1204 ± 0.0029
0.1153 ± 0.0023
0.1134 ± 0.0034
0.1196 ± 0.0013
0.1197 ± 0.0039
i◦
86.2 ± 0.5
86.2 ± 1.0
87.3 ± 0.6
86.9 ± 0.8
86.3 ± 0.5
86.7 ± 0.9
86.8 ± 0.8
87.4 ± 1.5
86.6 ± 0.3
85.9 ± 1.0
rA
rb
0.1218 ± 0.0020
0.1240 ± 0.0090
0.1135 ± 0.0039
0.1147 ± 0.0068
0.1216 ± 0.0044
0.1163 ± 0.0076
0.1171 ± 0.0065
0.1102 ± 0.0089
0.1148 ± 0.0026
0.1261 ± 0.0100
0.01496 ± 0.00086
0.0147 ± 0.0014
0.01333 ± 0.00060
0.0136 ± 0.0011
0.0147 ± 0.00068
0.0140 ± 0.0011
0.01349 ± 0.00095
0.0125 ± 0.0013
0.01373 ± 0.00042
0.0151 ± 0.0016
86.59 ± 0.18
0.1168 ± 0.0016
0.0139 ± 0.0002
Anderson et al. (2012)
0.1398
0.1260 ± 0.0030
86.02+1.11
−0.86
0.1242 ± 0.0102
0.0157
8
L. Mancini et al.
Table 5. Derived physical properties of WASP-44 from using each of five different theoretical stellar models. In each case gb = 21.5 ±
1.6 m s−2, ρA = 1.414 ± 0.058 ρ⊙ and Teq = 1304 ± 37 K.
This work
(Claret models)
This work
(Y2 models)
This work
This work
This work
(Teramo models)
(VRSS models)
(DSEP models)
Kb (km s−1)
MA (M⊙)
RA (R⊙)
log gA (cgs)
Mb (Mjup)
Rb (Rjup)
ρb (ρjup)
Θ
a (AU)
156.1 ± 4.1
0.968 ± 0.077
0.881 ± 0.025
4.534 ± 0.018
0.901 ± 0.075
1.020 ± 0.033
0.793 ± 0.070
151.7 ± 1.0
0.888 ± 0.018
0.856 ± 0.016
4.521 ± 0.010
0.851 ± 0.056
0.992 ± 0.019
0.816 ± 0.069
152.6 ± 4.2
0.904 ± 0.074
0.862 ± 0.024
4.524 ± 0.018
0.861 ± 0.073
0.997 ± 0.033
0.812 ± 0.072
152.7 ± 4.1
0.907 ± 0.073
0.862 ± 0.024
4.524 ± 0.018
0.863 ± 0.073
0.998 ± 0.032
0.811 ± 0.072
153.7 ± 3.2
0.924 ± 0.058
0.865 ± 0.020
4.530 ± 0.016
0.874 ± 0.067
0.994 ± 0.027
0.831 ± 0.071
0.0640 ± 0.0046
0.0658 ± 0.0045
0.0654 ± 0.0048
0.0654 ± 0.0047
0.0656 ± 0.0046
0.03508 ± 0.00093
0.03409 ± 0.00023
0.03429 ± 0.00093
0.03432 ± 0.00092
0.03454 ± 0.00072
Table 6. Final physical properties of the WASP-44 system. The
first errorbar for each parameter is the statistical error, which
stems from the measured spectroscopic and photometric parame-
ters. The second errorbar is the systematic error arising from the
use of theoretical stellar models, and is given only for those pa-
rameters which have a dependence on stellar theory. The results
from Anderson et al. (2012) are included for comparison.
This work (final)
Anderson et al. (2012)
MA (M⊙)
RA (R⊙)
log gA (cgs)
ρA (ρ⊙)
Mb (Mjup)
Rb (Rjup)
gb (ms−2)
ρb (ρjup)
Teq (K)
Θ
a (AU)
Age (Gyr)
0.917 ± 0.077 ± 0.051
0.865 ± 0.025 ± 0.016
4.526 ± 0.018 ± 0.008
1.414 ± 0.058
0.869 ± 0.075 ± 0.032
1.002 ± 0.033 ± 0.018
21.5 ± 1.6
0.808 ± 0.072 ± 0.015
1304 ± 37
0.0652 ± 0.0048 ± 0.0012
0.951 ± 0.034
0.927+0.068
−0.074
4.481+0.068
−0.057
1.19+0.32
−0.22
0.889 ± 0.062
1.14 ± 0.11
15.7+3.4
−3.0
0.61+0.23
−0.15
1343 ± 64
−
0.03445 ± 0.00093 ± 0.00063
0.03473 ± 0.00041
4.1+5.9 +4.1
−6.0 −2.4
−
tables7, a 0.875 MJ planet with 50 MEarth core at 0.045 AU
from the Sun (age 3.16 Gyr), is 0.991 RJup. On the contrary,
the expected radius for a core-free planet with the same
properties is 1.119 RJup.
3.4 Variation of planetary radius with wavelength
As an additional possibility for the GROND data, we made
an attempt to investigate the possible variation of the radius
of WASP-44 b with wavelength. This is quite a difficult task
because the relative faintness of the host star (V ∼ 12.9)
makes this system less than ideal for optical photometric
studies, especially from the ground. Moreover, as already
mentioned, our in-focus observing strategy has proved to
be ill-suited to obtaining high-precision photometry, partic-
ularly in the NIR bands. Finally, by using broad-band filters
we are forced to average our measurements of the planet ra-
dius in each band over a quite larger range of the wavelength
(the best case in the i′ filter where its width is "only" 100
7 The tables, interpolated from the giant planet thermal evolu-
tion models described in Fortney et al. (2007), are available at
http://www.ucolick.org/ jfortney/models.htm
nm). For these reasons, the results of this section should be
used with care.
The GROND instrument was conceived only for the
follow-up of gamma-ray bursts, and was not designed to al-
low filter changes for individual observing sequences. We
were therefore unable to use filters with narrow pass-
bands covering specific wavelength ranges. We note that
the four optical bands of GROND were already used by
de Mooij et al. (2012) to investigate the atmosphere of
GJ 1214 b, a 6.55 M⊕ transiting planet. Here we try for the
first time to use all the seven bands, exploiting the full po-
tential of GROND.
Following the same strategy used by Southworth et al.
(2012b) and Mancini et al. (2013), we proceeded as follows.
The two GROND datasets were combined by phase and ac-
cording to passband, and then fitted with all parameters
fixed to the final values given in Table 4, with the exception
of |
1602.02758 | 1 | 1602 | 2016-02-08T21:00:03 | In hot water: effects of temperature-dependent interiors on the radii of water-rich super-Earths | [
"astro-ph.EP"
] | Observational advancements are leading to increasingly precise measurements of super-Earth masses and radii. Such measurements are used in internal structure models to constrain interior compositions of super-Earths. It is now critically important to quantify the effect of various model assumptions on the predicted radii. In particular, models often neglect thermal effects, a choice justified by noting that the thermal expansion of a solid Earth-like planet is small. However, the thermal effects for water-rich interiors may be significant. We have systematically explored the extent to which thermal effects can influence the radii of water-rich super-Earths over a wide range of masses, surface temperatures, surface pressures and water mass fractions. We developed temperature-dependent internal structure models of water-rich super-Earths that include a comprehensive temperature-dependent water equation of state. We found that thermal effects induce significant changes in their radii. For example, for super-Earths with 10 per cent water by mass, the radius increases by up to 0.5$\,$R$_\oplus$ when the surface temperature is increased from 300 to 1000$\,$K, assuming a surface pressure of 100$\,$bar and an adiabatic temperature gradient in the water layer. The increase is even larger at lower surface pressures and/or higher surface temperatures, while changing the water fraction makes only a marginal difference. These effects are comparable to current super-Earth radial measurement errors, which can be better than 0.1$\,$R$_\oplus$. It is therefore important to ensure that the thermal behaviour of water is taken into account when interpreting super-Earth radii using internal structure models. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- 16 (2016)
Preprint 16 October 2018
Compiled using MNRAS LATEX style file v3.0
In hot water: effects of temperature-dependent interiors on
the radii of water-rich super-Earths
Scott W. Thomas & Nikku Madhusudhan
Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, United Kingdom
January 2016
ABSTRACT
Observational advancements are leading to increasingly precise measurements of super-
Earth masses and radii. Such measurements are used in internal structure models
to constrain interior compositions of super-Earths. It is now critically important to
quantify the effect of various model assumptions on the predicted radii. In particu-
lar, models often neglect thermal effects, a choice justified by noting that the thermal
expansion of a solid Earth-like planet is small. However, the thermal effects for water-
rich interiors may be significant. We have systematically explored the extent to which
thermal effects can influence the radii of water-rich super-Earths over a wide range of
masses, surface temperatures, surface pressures and water mass fractions. We devel-
oped temperature-dependent internal structure models of water-rich super-Earths that
include a comprehensive temperature-dependent water equation of state. We found
that thermal effects induce significant changes in their radii. For example, for super-
Earths with 10 per cent water by mass, the radius increases by up to 0.5 R⊕ when the
surface temperature is increased from 300 to 1000 K, assuming a surface pressure of
100 bar and an adiabatic temperature gradient in the water layer. The increase is even
larger at lower surface pressures and/or higher surface temperatures, while changing
the water fraction makes only a marginal difference. These effects are comparable to
current super-Earth radial measurement errors, which can be better than 0.1 R⊕. It
is therefore important to ensure that the thermal behaviour of water is taken into
account when interpreting super-Earth radii using internal structure models.
1 INTRODUCTION
One of the most interesting classes of planets today is
the class of super-Earths, planets with masses between 1
and 10 M⊕. With no analogues in the solar system, it is
not known whether they are scaled up rocky planets or
scaled down Neptunes. About 40 super-Earths with mea-
sured masses and radii are currently known. Their radii
range from 1 to 7 R⊕.1 With the potential to have mod-
erate atmospheres and plate tectonics, super-Earths repre-
sent an important class of planets in the broader context
of planetary diversity and planetary habitability (Haghigh-
ipour 2011; Baraffe et al. 2014).
Recent observational advancements are leading to in-
creasingly precise measurements of masses and radii of these
small planets. Such measurements are being used with inter-
nal structure models to place constraints on the interior com-
positions of super-Earths. Many planets are well-described
by multi-layer models consisting of iron, silicates, and water
(e.g. Valencia et al. 2006; Fortney et al. 2007; Sotin et al.
2007; Seager et al. 2007) and others have included layers
of hydrogen or other volatiles to explain the inflated radii
1 This number is taken from the exoplanets.org database (con-
firmed planets only).
c(cid:13) 2016 The Authors
of some super-Earths (e.g. Rogers & Seager 2010a; Lopez
et al. 2012). Given the high-precision radii measurements,
it is now critically important to quantify the dependence of
predicted radii of super-Earth models on the various model
assumptions.
Our goal
is to quantify the effects of temperature-
dependent internal structure models on the predicted radii
of super-Earths. An understanding of the effects of temper-
ature on the internal structures of planets is especially rel-
evant as our observational capabilities for measuring radii
and masses improve. In particular, we are interested in un-
derstanding to what degree the observable radius of a planet
may be affected by thermal expansion of its interior. We use
water as a case study for this, focusing on super-Earth plan-
ets consisting of a rocky Earth-like core underneath a heated
water layer.
The remainder of this section provides an overview of
how these planetary interior models can be useful, why we
expect temperature-dependent models to be different and
why water-rich planets make interesting test cases for as-
sessing whether this temperature dependence is significant.
2
Thomas & Madhusudhan
1.1 Planetary interior models
As atmospheric characterisation techniques improve, the
question of what lies beneath the atmosphere has naturally
arisen. We care about planetary interiors because they are
linked to the formation history of the planet, because they
are shaped by and shape the planetary atmosphere and be-
cause they are key to answering questions about habitability
(Sotin et al. 2010). Understanding these exoplanets also al-
lows us to place our own Earth into context: how unique
are we? We therefore seek to understand, if not the interi-
ors of individual exoplanets, at least something about broad
classes of planets. But it is here that we are confronted by
a lack of data, because we have very little ability to directly
probe the interiors of exoplanets.
This lack of a rich source of observational data for plan-
etary interiors means that we rely strongly on models. Even
inside our solar system, our knowledge of planetary interi-
ors is limited by the indirect ways in which we can probe
them. On Earth we have the advantage of seismic measure-
ments, and in our solar system we have various gravitational
moments to constrain interior structures. Outside the solar
system we have only the masses and radii of planets to work
with. Models from first principles (numerical or analytical
models based on the physics of solid and liquid spheres)
therefore dominate the field.
Planetary interior models are a worthwhile starting
point to make sense of the limited observational data we
have. These models are inspired by earlier successes with
stellar structure models, which are key to interpreting ob-
servations of stars. Others had previously considered the in-
ternal structures of planets in our solar system (for example,
Hubbard & MacFarlane 1980), but the study often taken as
the base for planetary interior modelling is by Zapolsky &
Salpeter (1969) who constructed mass -- radius relations for
large homogeneous isothermal spheres. A number of internal
structure models have been developed for exoplanets, start-
ing with early works a decade ago (e.g. Valencia et al. 2006;
Fortney et al. 2007; Sotin et al. 2007; Seager et al. 2007).
The ever-increasing number of known exoplanets, many of
which have both mass and radius measurements, are a di-
verse and interesting set of objects to which to apply these
models.
The first way in which planetary interior models can be
useful is to make broad inferences about the structure of a
planet. There is some information available about any planet
despite an inherent degeneracy between different composi-
tions. We can immediately exclude certain classes of mod-
els: for example, small planets with large radii must almost
certainly have large hydrogen envelopes. We can also take
more sophisticated approaches. Sotin et al. (2007) modelled
planets by fixing their compositions based on the properties
of the host star. Madhusudhan et al. (2012) argued for a
carbon-rich interior in the exoplanet 55 Cnc e based on its
carbon abundance and on its density matching that of pure
carbon. Dorn et al. (2015) also showed that mass and radius
alone can constrain the size of a planet's core if we assume
it is pure iron.
We can also hope to make progress in a statistical sense
by examining populations of planets. Such progress is possi-
ble even if we are unable to pin down the precise structure
of an individual planet. There are promising advances in
this direction already. These usually involve inverse Bayesian
analyses. For example, Rogers (2015) investigated the size
demographics of planetary populations and set an approxi-
mate boundary of 1.6 R⊕ beyond which planets are likely to
have gaseous envelopes.
Finally, interior structure models may be useful when
combined with prescriptions for planetary formation. Mor-
dasini et al. (2012) took this approach, combining inte-
rior structure calculations with models of the protoplane-
tary disk to produce synthetic populations of planets. Lopez
et al. (2012) also made model planets and explored how they
evolve and lose mass through time (see also Owen & Wu
2015).
If we are to use mass and radius to constrain the inte-
rior structure of a planet, we should ensure that our models
are precise and accurate. But more importantly, we should
understand where our models need to be precise and accu-
rate and where such effort is wasted. We therefore require a
thorough understanding of what factors can affect the mass --
radius relation. We also need to know to what extent we are
able to invert the relation to determine a composition.
The internal structure of a planet is not well-constrained
by its mass and radius alone (Rogers & Seager 2010a). How-
ever, we know that we can obtain some compositional con-
straints from observations of the planet and its host star.
Above, we mentioned works by Sotin et al. (2007) and Mad-
husudhan et al. (2012), who used host star information in
this way. Dorn et al. (2015) also used probabilistic models,
incorporating the host star chemical abundances, to con-
clude that "uncertainties on mass, radius, and stellar abun-
dance constraints appear to be equally important." Grasset
et al. (2009) indicated the need for good radius measure-
ments, especially for dry silicate-rich planets for which nu-
merical models can provide radius estimates to precisions of
less than 5 per cent. Unterborn et al. (2015) used a mineral
physics toolkit to perform a sensitivity analysis for rocky
super-Earths, concluding that the mass -- radius relationship
is most strongly altered by the core radius and the presence
of light elements in the core.
The presence of an atmosphere could also contribute sig-
nificantly to the observed radius. Rogers & Seager (2010b)
have modelled isothermal super-Earth interiors overlaid by a
volatile atmosphere. Additionally, Valencia et al. (2013) con-
sidered coupled atmosphere -- interior models, which also in-
cluded atmospheric mass loss, and explored the dependence
of radii on various model parameters such as the irradiation,
water content and metallicity. The effect of an atmosphere
is important, especially given that observations can probe
spectral ranges where atmospheric absorption could be sig-
nificant (Madhusudhan & Redfield 2015).
Though the factors above are all important, the effect
of temperature on the mass -- radius relation has not been
thoroughly explored (but see e.g. Valencia et al. 2013). This
is for several reasons. First, its effects are thought to be
relatively minor in the first place: Howe et al. (2014) esti-
mate that the effect of thermal corrections on an iron-silicate
planet's radius is approximately 5 per cent. Grasset et al.
(2009) also describe how the radius of an Earth-like planet
is not strongly affected by temperature changes. If the effect
is small compared with current observational uncertainties,
it is not necessarily relevant. Secondly, modelling is easier if
we assume zero-temperature or isothermal spheres of mate-
MNRAS 000, 1 -- 16 (2016)
Thermal effects in super-Earth interiors
3
the phase changes throughout. Zeng & Sasselov (2014) chose
to explore evolutionary effects, following the phase transi-
tions within model water-rich planets. They comment that
"[phase] transformations may have a significant effect on the
interior convective pattern and also the magnetic field of
such a planet, but they may only affect the overall radius
slightly." Our present study addresses the question of ex-
actly how much temperature variations affect the structure
and radius of water-rich planets and whether such effects are
observable.
2 METHOD
Guided by the motivations above, we quantified the thermal
effects in super-Earths which contain significant amounts
of water. By thermal effects we mean three effects in par-
ticular. First, we mean the thermal expansion of a heated
water layer on the surface or within the planet's interior:
this contrasts with models which treat the planets as cold
spheres. Secondly, we mean any temperature gradient estab-
lished within the planet: this is in contrast to the isothermal
case. And thirdly, we expect phase transitions within the
water layer if the temperature and pressure cross one of the
boundaries between different phases of water seen in Fig. 1.
To quantify these effects required several steps. First,
we selected an appropriate temperature-dependent equation
of state. We built planetary interior structure models that
included this equation of state. We incorporated a realistic
temperature gradient into these models. Finally, we explored
the model parameter space. In particular, we compared the
mass -- radius relationships for these water worlds across a
range of surface pressures, surface temperatures and interior
compositions.
In this section, we first explain how we built mod-
els of super-Earth interiors: we constructed layered one-
dimensional models of water, rock and iron in varying pro-
portions. We discuss our approach to the temperature struc-
ture: we treated the temperature gradient as adiabatic and
self-consistently calculated it from the equation of state, an
approach which naturally handles phase boundaries within
the water layer. We explain how we used these models to
construct temperature-dependent mass -- radius relations for
both homogeneous and layered planets. Next we present the
equation of state for water that we used. It is comprehen-
sive over the pressure and temperature range relevant to
super-Earth interiors. We highlight the difficulty of dealing
with disparate sources of experimental and theoretical data
in different phases, as well as sources of uncertainty within
the equation of state. Finally, we present comparisons with
previous works to verify that our structural modelling code
works appropriately.
2.1 Interior structure modelling
We constructed temperature-dependent planetary interior
models of water-rich super-Earths. We considered the planet
to be spherically symmetric, non-rotating and non-magnetic.
The following equations govern the structure of such a plan-
etary interior.
Figure 1. Phase diagram of water. Water has a rich and interest-
ing phase structure. Here we show some of the key phases which
are relevant when modelling a watery planet:
liquid, vapour,
and solid ice Ih, but also more exotic phases such as the high-
pressure ices. Lines mark the boundaries of each phase as given
by Choukroun & Grasset (2007) and Wagner & Pruss (2002).
rial, because we do not have to deal with energy transport
within the planet. Finally, the data on thermal expansion
of heavy elements are sparse at the high temperatures and
pressures characteristic of planetary interiors (Baraffe et al.
2008). Therefore mass -- radius relations or models of individ-
ual planets traditionally had no temperature dependence at
all (Zapolsky & Salpeter 1969; Seager et al. 2007) but it is
increasingly being included and thermal effects on radii are
being explored (for example, see Valencia et al. 2013).
1.2 Temperature dependence of water-rich
planets
The degree to which thermal structure may affect the prop-
erties of a water-rich planet has not yet been well studied.
Super-Earth planets with significant water layers, sometimes
called waterworlds, provide an interesting testbed for our in-
vestigation. They may display more significant variation in
their properties, both observable and internal, than purely
Earth-like (iron and silicate) planets. They are therefore a
worthwhile target for this study.
Water presents an opportunity to assess thermal effects
in a material that has a rich and interesting phase struc-
ture across a large temperature and pressure range (Fig.
1). At low temperature and pressure, water exists as a liq-
uid, vapour, or solid (Ice Ih). At high pressure, it takes on
a number of alternate ice forms (Ice V, VI, VII, X, etc.)
(Choukroun & Grasset 2007). It can also exist as a low-
density supercritical fluid or superheated vapour. This all
means that the behaviour of water layers is thermally inter-
esting. The behaviour of water is also strongly linked to ques-
tions of habitability because Earth-sized solid planets with
oceans provide the best approximation to the one planet
known to harbour life.
Others have previously investigated the structures of
planets containing a significant water component. For ex-
ample, Ehrenreich et al. (2006) studied the internal struc-
ture of the exoplanet OGLE 2005-BLG-390Lb, modelling
MNRAS 000, 1 -- 16 (2016)
10510610710810910101011Water pressure / Pa102103Water temperature / KLiquidVapourSupercriticalfluid/plasmaIce IhVIVIIXCritical point4
Thomas & Madhusudhan
2.1.1 Planetary structure equations
The mass continuity equation,
dr
dm =
1
4πr2ρ ,
(1)
links r, the radius of a spherical shell, to the mass m interior
to the shell and the density ρ of the shell. The equation of
hydrostatic equilibrium,
our equation of state section. Equation 4 combined with
equation 1 gives the temperature gradient in terms of the
mass co-ordinate,
dT
dm = − TαGm
ρcp4πr4 .
(6)
dP
dm = − Gm
4πr4 ,
(2)
2.1.2 Solving the structural equations
where P is the pressure at the shell and G is the gravita-
tional constant, ensures a balance of pressure and gravity.
The equation of state,
ρ = ρ(P,T ),
(3)
is used to calculate the density of the material in question
from its pressure and temperature T .
Previous models of super-Earth interiors have treated
temperature gradients within the planet in a number of
ways. From simple to complex, these include:
1. Isothermal models, which use only the equations above
and an isothermal equation of state of the form ρ = ρ(P).
This is the approach taken by Seager et al. (2007).
2. Simple temperature prescriptions:
a) One may assume a fixed temperature -- pressure re-
lation T = T (P) so as to reduce the equation of state to
the form ρ = ρ(P) and then use the equations above. For
example, Zeng & Sasselov (2013) chose the melting curve
of water for this purpose.
b) Or one may choose a temperature profile T = T (r)
for the planet (perhaps scaled appropriately to an inter-
nal or external boundary temperature) and then use the
equations above.
3. An adiabatic or conductive temperature gradient or
some combination of the two. For example, Valencia et al.
(2010) used a convective interior with conductive boundary
layers.
4. A full treatment, which adds an energy transport equa-
tion to the equations above then self-consistently solves this
with a prescription for luminosity. For example, Wagner
et al. (2011) modelled an adiabatic core underneath a ra-
diogenically heated mantle.
We did not explicitly handle energy transport in the
manner of the fourth option. Instead we chose the third ap-
proach and assumed an adiabatic (isentropic) temperature
gradient throughout the planet. The equation for the adi-
abatic temperature gradient, as given by Milone & Wilson
(2014), is
dT
dr = − Tαg
cp
.
(4)
where g = Gm/r2 is the gravity at the shell, cp is the isobaric
heat capacity and α is the volumetric thermal expansion co-
efficient. This is sometimes written αV and is defined as the
fractional increase in volume per unit temperature increase,
α =
1
V
∂V
∂ T
= − ∂ lnρ
∂ lnT
.
(5)
(cid:12)(cid:12)(cid:12)(cid:12)p
(cid:12)(cid:12)(cid:12)(cid:12)p
Our sources for these latter two coefficients are detailed in
Together, equations 1, 2, 3 and 6 define a structural model:
three ordinary differential equations and an equation of state
linking pressure, temperature and density. The choice of how
to solve this system depends on one's aim. A common ap-
proach has been to treat it as a boundary value problem;
that is, to integrate the structural equations from initial
conditions at the surface or centre of the planet. For exam-
ple, Seager et al. (2007) approached the isothermal problem
(equations 1, 2 and 3 only) from the inside out, choosing
appropriate central pressures at the (r = 0,m = 0) boundary
and building their models outward from there. We instead
integrate from the outside in, an approach taken by several
others (e.g. Rogers & Seager 2010a; Madhusudhan et al.
2012). This has the advantage of allowing us to specify the
surface temperature and pressure as boundary conditions.
These surface boundary conditions are more closely linked
to observable parameters than the central pressure and tem-
perature.
We used a Lagrangian system, where the mass interior
to a given shell is the independent variable; this is reflected
in equations 1, 2 and 3. It is in contrast to the Eulerian co-
ordinate system used by Seager et al. (2007), who take the
radius r as the independent variable. Rogers & Seager (2012)
claim that this formulation is more stable under numerical
integration. We also found it more convenient to be able
to specify differentiated planets in terms of mass fractions
rather than radial distances.
We solved this boundary value problem using a shooting
method2. This method used a series of trial solutions, adjust-
ing the initial conditions as necessary based on the difference
between the expected and actual values at the end of the in-
tegration domain. For the initial trial solution, we specified
the surface boundary conditions: total planetary mass M,
surface pressure P(M), and surface temperature T (M). We
also specified a search bracket for the radius [R1(M),R2(M)].
Our code3 used a fixed-step fourth-order Runge -- Kutta in-
tegrator to solve the system of differential equations above.
In each successive trial, it iteratively adjusted the radius
boundary condition R(M) according to the bisection root-
finding method to ensure that the radius approached zero
as the mass approached zero. We further required that r re-
mained positive: this avoided any numerical difficulty arising
from the behaviour of the equations at r = 0. We deemed the
system to be converged acceptably when the central radius
r(m = 0) was between 0 and 100 m.
2 For an example implementation, see Press (2007), chapter 18.
3 Our code was written in Julia.
MNRAS 000, 1 -- 16 (2016)
2.1.3 Mass -- radius relations with thermal expansion and
2.2.1 Previous equations of state for water
Thermal effects in super-Earth interiors
5
multiple layers
We used our models to produce mass -- radius relations for ho-
mogeneous spheres of water as well as differentiated multi-
layer models. We did this first for the homogeneous isother-
mal case (in the vein of Zapolsky & Salpeter 1969) and then
extended our models to include an adiabatic temperature
gradient. Our differentiated multi-layer models included a
water layer on top of a silicate mantle and an iron core. To
do this, they treated the equation of state, equation 3, as
piecewise in the mass co-ordinate. For example, consider a
model which has a 5 per cent (by mass) water layer on top
of a silicate mantle. For this model, we begin by evaluating
equation 3 using the water equation of state. We then switch
to the silicate equation of state once m, the mass interior to
the spherical shell in equations 1, 2 and 6, drops below 95
per cent of the planetary mass. It is possible to choose the
integration grid such that this occurs exactly at the end of
an integration step. However, in practice a sufficiently fine
grid is also acceptable.
We ignored thermal effects within the iron and silicate
layers. The effect of thermal expansion in these solids is
thought to be low (see Seager et al. 2007; Grasset et al.
2009). Including the expansion effects of these materials
would be very simple, but we have not yet collated the equa-
tion of state data which would enable us to do so. Because we
ignored thermal expansion in these layers, we modelled them
as isothermal, which follows from setting α = 0 in equation
6 so that dT /dm = 0.
We aimed to accurately capture the density change of
water at its phase boundaries. Our equation of state for wa-
ter therefore included its phase transitions, which appear as
density discontinuities in pressure -- temperature space. When
calculating the adiabatic temperature profile, we enforced
temperature and pressure continuity at these phase bound-
aries. We did this by ensuring that the equation for the adi-
abatic temperature gradient, equation 6, was finite and con-
tinuous. This effectively split the adiabatic temperature pro-
file into several different sections, consisting of one separate
adiabat for each phase and meeting at the phase boundaries
of water. By handling each phase separately, we avoided the
numerical difficulty of taking a derivative (equation 5) across
a density discontinuity.
We note that it is possible to fix the radius and let
another parameter vary instead. This could be the mass,
surface temperature, surface pressure or the position of a
layer boundary within the planet. Other studies have used
this approach to infer potential compositions for planets of
known mass and radius. We instead left the radius free to
investigate how much it was affected by the water layer's
thermal expansion. This was the primary goal of this study.
2.2 Equation of state
As the goal of this study was to investigate thermal ef-
fects, we required a temperature-dependent equation of state
for water. This allowed us to treat thermal expansion self-
consistently in our models. We synthesized an equation of
state for water from the best available experimental and the-
oretical data over a wide range of pressure and temperature.
MNRAS 000, 1 -- 16 (2016)
There was no single comprehensive equation of state data
set available for water over the entire pressure and tempera-
ture range relevant to super-Earth interiors. Previous studies
have approached this problem by stitching together equa-
tions of state which are valid for different pressure regimes.
For example, Seager et al. (2007) took this approach with
water, combining three temperature-independent equations
of state for ice VII:
• the Birch -- Murnaghan equation of state at low pres-
sures,• density functional theory calculations at intermediate
• the Thomas -- Fermi -- Dirac model at very high pressures.
pressures and
The piecewise function defined in this way is appropri-
ate across a wide pressure range.
This pressure piecewise approach neglects temperature
dependence in the equation of state but provides a robust
approximation that is easy to evaluate. In some cases, stitch-
ing the data in this fashion has revealed that a simpler func-
tional form works just as well. For example, this is the case
in the "polytropic equation of state" used by Seager et al.
(2007). Such simple functional forms for the equation of state
have been used successfully to model planets as cold spheres
since the work of Zapolsky & Salpeter (1969). In other cases,
a more detailed functional form is needed to capture the be-
haviour of the material fully; this is especially true if it un-
dergoes phase transitions. For example, the IAPWS formula-
tion described by Wagner & Pruss (2002) uses a complicated
series of equations fitted to various sources of experimental
data for the behaviour of water in the vapour and liquid
phases.
In choosing equations of state, previous authors have
taken similar approaches to the stitching technique above.
Although the choice of the exact equations has varied as
new experimental data were released, few of these studies
included thermal expansion. Howe et al. (2014) provided a
comprehensive overview of the equations of state chosen in
previous works to model planetary interiors. They included
several different materials of interest for planetary interiors:
water ice, iron, and silicates. We repeated this exercise, fo-
cusing exclusively on the water equations of state across all
its phases. Table 1 summarises our findings.4
2.2.2 Our equation of state
We extended the piecewise approach described above to
include temperature as a second dimension in parameter
space. We drew from a number of the sources of data listed
in Table 1. Our stitched equation of state is valid over a wide
domain: its temperature domain is from 275 K to 24000 K,
and its pressure domain is from 105 Pa (1 bar) upwards. Our
approach was similar to that of Senft & Stewart (2008), who
generated a "5-Phase" equation of state across different liq-
uid, vapour, and ice phases. However, their work focused on
the lower temperatures needed to model impact craters. We
4 Where abbreviations are used in this table and Table 2, Table
3 indicates from which studies they come.
6
Thomas & Madhusudhan
Table 1. Previous studies on planetary interior structures use a
variety of equations of state for water.
Work(s)
Water equation of state used
Table 2. We used a variety of equations of state in our final
models. "Tabular" indicates that we interpolated between values
specified in the paper. "Functional" indicates that we used the
functional form given in the paper.
TFD, BME, MGD
Equation of state
Type
Region of validity
Baraffe et al. (2008);
Baraffe et al. (2014)
Fortney et al. (2007)
Simple power law from Hubbard &
MacFarlane (1980)
H2O-REOS
Fortney &
Nettelmann (2009)
Grasset et al. (2009) MGD; Vinet; BME; TFD; ANEOS;
Guillot (1999)
Howe et al. (2014)
Hubbard &
MacFarlane (1980)
Hubbard & Marley
(1989)
Lopez et al. (2012)
Madhusudhan
(2012)
More et al. (1988)
Nettelmann et al.
(2008)
Nettelmann et al.
(2011)
Redmer et al. (2011)
Rogers & Seager
(2010b)
Seager et al. (2007);
Rogers & Seager
(2010a); Zeng &
Sasselov (2014)
Senft & Stewart
(2008)
Sotin et al. (2007);
Sotin et al. (2010)
Valencia et al.
(2006)
Valencia et al.
(2010)
Vazan et al. (2013)
Wilson & Militzer
(2012); Wilson et al.
(2013)
Zeng & Sasselov
(2013)
Belonoshko & Saxena (1991)
Hubbard & Marley (1989)
Vinet
Simple power law
Exponential polynomial EOS without
temperature dependence
H2O-REOS
BME
Quotidian EOS (ion EOS with
Thomas -- Fermi model)
LM-REOS
H2O-REOS
French et al. (2009)
IAPWS; IAPWS extrapolations; TFD
Low-temperature polytropic EOS
IAPWS; Feistel & Wagner (2006);
Stewart & Ahrens (2005); BME
BME with thermal expansion (MGD)
BME with thermal expansion
French et al. (2009); SESAME
Quotidian EOS; TFD
DFT
Frank et al. (2004); French et al.
(2009); TFD
have explicitly included much higher temperatures so as to
capture the behaviour of large super-Earth planets: we ex-
pect the cores of these to reach thousands of Kelvin. Valen-
cia et al. (2010) also constructed a similar equation of state,
though using only data from SESAME and the IAPWS for-
mulation.
We endeavoured to choose equations of state that were
most representative of the thermal behaviour of water across
this temperature and pressure domain. We were guided by
two principles in doing so. First, as demonstrated in Fig. 3,
we expect thermal expansion effects to approach zero as the
pressure increases: this is a consequence of the equations
of state approaching the high-pressure TFD limit. There
are significant temperature effects at lower pressures, and
it is these effects we expected to be most important in our
IAPWS
Tabular
French et al.
(2009)
Tabular
Feistel & Wagner
(2006)
Sugimura et al.
(2010)
Vinet + MGD
correction using
parameters from
Fei et al. (1993)
TFD
Seager et al.
(2007)
Choukroun &
Grasset (2007)
IAPWS
extrapolations
Vapour and liquid phases; 0.05
to 1000 MPa and 252.462 to
1273 K
Superionic, plasma and
high-pressure ice phases; 79 to
9.87× 106 MPa and 1000 to
24000 K. We did not use table
VIII from this work, as this
low-density data disagrees with
the IAPWS formulation.
Ice Ih; 0 to 200 MPa and 0 to
273 K
Ice VII; 18880 to 50250 MPa
and 431 to 881 K
Ice VII
Tabular
Tabular
Functional
Functional
Ice X
Functional
Ice VIII -- X transition
Functional
Ices I, III, V, VI; phase
boundaries as specified by
Dunaeva et al. (2010)
Functional Remaining regions
study. Secondly, we aimed for a full treatment of density
changes over phase boundaries. Accordingly, we used the
phase boundaries specified by Dunaeva et al. (2010) to di-
vide the temperature -- pressure phase space into regions cor-
responding to different phases of water. We then chose an
appropriate equation of state to represent each phase.
Our equation of state is for pure water only. Others have
investigated how impurities may affect the equation of state
and the planet's properties. For example, Levi et al. (2014)
included a methane component in their models, resulting in
a new phase of water (filled ice) which changes the planet's
thermal profile. They note that, while neglecting volatiles is
an impediment to understanding the planet's atmosphere,
pure water models may be sufficient for planetary mass --
radius relations.
In selecting the equations of state we were often faced
with choices between different sources of data. The exact be-
haviour of water at very high pressures is still uncertain and
experimental and theoretical results are sometimes in con-
flict (Baraffe et al. 2014). Ensuring absolute accuracy of the
chosen equations of state was therefore a secondary priority.
In general, we preferred more recent data to older data, we
prioritised measured and tabulated values over functional
approximations, and we chose representations that included
temperature dependence over those that did not. In the fol-
lowing paragraphs, we describe our equation of state choices
and summarise them in Table 2.
Liquid and vapour: The behaviour of water in the liquid
and vapour phases is well understood and there are plenty
MNRAS 000, 1 -- 16 (2016)
Thermal effects in super-Earth interiors
7
Figure 2. Phases and data sources for our water equation of state. Our equation of state covers a wide range of temperature -- pressure
space. On the left, we show some of the our key data sources we used, and their regions of validity: the IAPWS formulation by Wagner
& Pruss (2002); the theoretical calculations of French et al. (2009); the piecewise equation of state described by Seager et al. (2007); the
Mie-Gruneisen-Debye (MGD) thermal correction approach for ice VII in Sotin et al. (2007); and the measurements of Sugimura et al.
(2010), which cover a small region of ice VII. We also show the relevant phase boundaries. On the right, we show the density variation
across the entire pressure -- temperature range. The density of water is more strongly affected by pressure across the range we consider,
but temperature also affects its density too, especially across the liquid -- vapour phase boundary and in the supercritical region.
nationals. Instead, to represent water liquid and vapour, we
selected the IAPWS (International Association for the Prop-
erties of Water and Steam) formulation (Wagner & Pruss
2002), which provides both tabular and functional data for
water in these phases. These are well-tested and validated by
years of experiments. Wagner & Pruss (2002) also claim that
the functional forms can be extrapolated outside the range
of the tables. We implemented the functional relationships
between temperature, density and pressure. Where appro-
priate, we numerically inverted these to give a relation of
the form ρ = ρ(P,T ). We then tested these against the ta-
bles to verify that we had replicated them correctly.
Ice VII: We explicitly chose a temperature-dependent for-
mulation because we expected ice VII to form a significant
fraction of the planet in the cases where the water layer
is large. This temperature-dependent formulation is in con-
trast to other studies which have assumed that the ice VII
layer is isothermal: for example, Rogers & Seager (2010b)
assumed no expansion in all solid layers, choosing to include
temperature effects only in the gaseous and liquid phases.
The best temperature-dependent formulation we found
for ice VII was the Mie-Gruneisen-Debye (MGD) thermal
correction approach described by Sotin et al. (2007). We
used a Vinet equation of state with this thermal correction,
taking the coefficients of Fei et al. (1993), within the ice
VII region delimited by the phase boundaries of Dunaeva
et al. (2010). However, we preferred the more recent tabu-
lated measurements of Sugimura et al. (2010) wherever these
were applicable; these are shown within the ice VII region
in Fig. 2.
Figure 3. Comparison of our equation of state with the high-
pressure limit. The TFD (Thomas -- Fermi -- Dirac) equation of state
is increasingly accurate in the high-pressure limit, where tempera-
ture effects on the water density disappear. We also show temper-
ature contours of our water equation of state. The TFD, which
has no temperature correction, is a poor approximation of the
behaviour of water at low pressures, especially across the liquid --
vapour phase boundary (vertical lines). But all other choices of
equation of state approach the TFD at high pressures, and so it
is appropriate in the TPa region and beyond.
of data available. We were unable to gain access to the
SESAME tables of Lyon & Johnson (1992) because there
are restrictions on the distribution of this data to non-US
MNRAS 000, 1 -- 16 (2016)
1051061071081091010101110121013103104IAPWSFrenchSeagerMGDSugimura10510610710810910101011101210131014LiquidVapourSupercriticalfluid/plasmaVIIXTheoreticalhigh-pressureicesCritical point100101102103104Water density / kgm−3Water pressure / PaWater temperature / K1051061071081091010101110121013Water pressure / Pa101102103104Water density / kgm−3TFD300 K500 K1000 K10 000 K8
Thomas & Madhusudhan
Supercritical fluid and plasma: French et al. (2009) pre-
sented quantum molecular dynamics simulations of high-
temperature and high-pressure plasma, ice, and superionic
fluid phases of water. We used their tables in the region be-
yond 1000 K and 1.86×109 Pa. Lopez et al. (2012) noted that
this region has recently been probed by laboratory experi-
ments thanks to Knudson et al. (2012), who strongly ad-
vocate "that [the French equation of state] be the standard
in modeling water in Neptune, Uranus, and 'hot Neptune'
exoplanets.". These temperatures and pressures are also rel-
evant to the interiors of super-Earths. We did not use the
low-density tables that they presented separately because
these differ significantly from the IAPWS results in the same
temperature and pressure range.
Low-temperature ices: For completeness, our equation
of state includes low-pressure ice Ih from Feistel & Wag-
ner (2006) as well as higher-pressure ices such as ice III, V
and VI. We took the phase boundaries from Dunaeva et al.
(2010) and used the temperature-dependent formulations for
these ices by Choukroun & Grasset (2007).
Ice X and beyond: We adopted the piecewise equation
of state of Seager et al. (2007) to describe the transition
from ice VII to ice X and beyond. This does not include any
temperature dependence: any interesting phase behaviour
of ice at these high pressures is increasingly theoretical and
unconfirmed by experiment. Temperature effects approach
zero at these high pressures anyway (Fig. 3), so we used
the Thomas -- Fermi -- Dirac equation of state for all regions
beyond 7686GPa which were not covered by one of the other
regions listed above.
Other regions: Finally, we filled in all other regions ac-
cording to the IAPWS formulation or extrapolations thereof.
In practice, the only regions not covered above were low-
pressure and high-temperature vapour regions, which we do
not expect to be relevant for our super-Earth interior mod-
els.
2.2.3 Dealing with fragmented data
We made no attempt to smooth or otherwise interpolate
between the different sources of data described above. This
approach means that sharp density changes across phase
boundaries are well-represented in the final equation of state.
This is desirable so that we may examine the differentiation
that results solely from phase transitions. It also results in
some artificial density discontinuities at the boundaries be-
tween different data sets. We believe that this has not af-
fected the results: these discontinuities are minor compared
with the density variations within each phase of water and,
in most cases, we also bounded the domain of each data set
to that of a particular phase.
Because we used disparate sources of data, we evaluated
the density at a given temperature and pressure in different
ways depending on the data source. Although we did not
smooth or interpolate between data sets, we needed to inter-
polate some data sources within the data set. Where data
were published in tabulated form on a structured grid, we
Table 3. Sources for the abbreviated equation of state designa-
tions used in this paper.
Equation of
state
Source
ANEOS
BME
DFT
H2O-REOS
IAPWS
LM-REOS
MGD
SESAME
TFD
Vinet
Thompson & Lauson (1972)
Birch -- Murnaghan equation of state; see
Poirier (2000)
Density functional theory; refers to
theoretical calculations which multiple
authors have performed
Nettelmann et al. (2011); includes IAPWS,
SESAME, French et al. (2009), Feistel &
Wagner (2006)
Wagner & Pruss (2002)
Nettelmann et al. (2008) (precursor to
H2O-REOS)
Mie-Gruneisen-Debye thermal pressure
expansion; described in Sotin et al. (2007)
Lyon & Johnson (1992)
Thomas -- Fermi -- Dirac; described in Salpeter
& Zapolsky (1967)
Vinet et al. (1987)
used simple two-dimensional linear interpolation5 to evalu-
ate the equation of state at points not lying on the grid.
Where data were published as unstructured points, we used
barycentric interpolation on the mesh of Delaunay triangles6
defined by these points. We also used this Delaunay mesh
to determine if a given (P,T ) pair lay within the domain of
a particular equation of state, allowing us to fall back to an-
other equation of state if necessary. We evaluated functional
forms of the equation of state as is, defining their domain by
means of a bounding box or a polygon in (P,T ) space taken
from the phase boundaries of Dunaeva et al. (2010).
Some of the equations of state used in this final syn-
thesized version were much simpler than others. This meant
that the evaluation time varied from point to point, from
very quick table lookups and interpolation to the slower
IAPWS formulae. In addition, any equation of state that
was specified in the inverse form P = P(ρ,T ) needed to be
numerically inverted to give the canonical form ρ = ρ(P,T )
used in our models. To avoid duplicating this calculation un-
necessarily, we re-sampled the final equation of state on to a
256 by 256 pressure -- temperature grid. Pre-computing and
tabulating the data in this way saved significant time. In our
trials, the resolution of the grid barely altered the proper-
ties of the planetary models. This suggests that the density
behaviour within a single phase region was more important
than any effects across phase boundaries that might be lost
by sampling from this discrete grid.
The equation of state we used necessarily has some un-
certainty in it, especially in regions near the critical point
of water (Wagner & Pruss 2002) and at high temperatures
and pressures where there are sometimes conflicting exper-
imental and theoretical data (Baraffe et al. 2008). The er-
ror in the equation of state varies depending on the original
data source. For the region encompassed by the IAPWS data
5 For multidimensional linear interpolation we used Dierckx.jl.
6 To construct a Delaunay tessellation we used VoronoiDelau-
nay.jl.
MNRAS 000, 1 -- 16 (2016)
(Wagner & Pruss 2002), the density uncertainty is approxi-
mately 0.01 per cent (liquid and solid), 0.03 to 0.1 per cent
(vapour), and up to 0.5 per cent in the region around and be-
yond the critical point. They give a more detailed breakdown
of these errors in their section 6.3.2, in particular fig. 6.1.
We estimate that the error beyond these regions is closer to
1 per cent if we extrapolate beyond the table and assume
that the uncertainty continues to increase at higher tem-
peratures and pressures. For the supercritical fluid, plasma
and superionic phases in the data of French et al. (2009),
they state that "the QMD EOS is accurate up to 1 per cent
for the conditions relevant for the giant planet's interiors of
our solar system." For the ice VII phase, the measurements
of Sugimura et al. (2010) have errors of between 0.003 per
cent and 0.5 per cent. Finally, it is not possible to give a
meaningful uncertainty estimate at higher pressures where
no measurements exist, but we do not treat the temperature
dependence there anyway.
2.2.4 Thermal expansion and heat capacity
Equation 6 requires both a heat capacity cp and a thermal
expansion coefficient α (defined in equation 5). Following
our goal of handling temperature effects appropriately, we
explicitly sought out temperature-dependent forms for these.
We used the IAPWS tables for heat capacity in the
liquid -- vapour range, then took the nearest available data
point from these tables for all other pressure -- temperature
points. This is because we could not find readily available
heat capacity data across the full range of phases in our
equation of state. This approach therefore does not reflect
any changes in heat capacity between the high-pressure ice
phases. The most significant effect is the change in heat ca-
pacity across the liquid -- vapour phase boundary, which we
do capture in our models.
We drew the thermal expansion coefficient α directly
from the equation of state by evaluating equation 5. We
used automatic differentiation7 where possible to evaluate
the derivative. In some cases this was not possible8 so we
used finite differencing9. As well as pre-computing the equa-
tion of state itself, we pre-computed and tabulated the ther-
mal expansion coefficient on the same pressure -- temperature
grid. Some previous works have assumed a fixed thermal
expansion coefficient: for example, Ehrenreich et al. (2006)
took a fixed value for α in their models. We believe that
our approach is more appropriate for understanding how
the temperature gradient and physical properties of a wa-
tery planet are affected by the thermal properties of water.
7 We used forward-mode automatic differentiation provided by
ForwardDiff.jl.
8 The Delaunay triangulation method in the library we used in-
corporates a method called floating-point filtering, which relies on
the specific properties of floating point numbers. It could not be
used with the automatic differentiation approach we used, which
evaluates functions as usual but replaces the inputs with a special
numeric type.
9 We used the package Calculus.jl.
MNRAS 000, 1 -- 16 (2016)
Thermal effects in super-Earth interiors
9
Figure 4. Validation of isothermal models. Our structural models
exactly reproduce previous results in the isothermal case. Here we
show mass -- radius relations for homogeneous isothermal spheres.
If we adopt identical equations of state to those used by Seager
et al. (2007), we obtain the same result. This serves as a verifica-
tion that we are correctly solving the structural equations. These
models used zero surface pressure and have no temperature de-
pendence: the equations of state are isothermal and are taken at
300 K.
2.3 Model verification
We verified our models by making mass -- radius diagrams as
described in the previous section and comparing them with
previous work.
2.3.1 The isothermal case
We checked that our models work in the isothermal case by
replicating the mass -- radius relations of Seager et al. (2007).
We exactly reproduced the mass -- radius relations when we
constructed homogeneous isothermal 300 K planets using the
equations of state specified in their paper, as shown in Fig.
4. We set the surface pressure of our models to zero, follow-
ing the boundary condition they used. The surface pressure
hardly affects the results because the equations of state are
for the solid phase only. This identical mass -- radius relation
verified that our integrator works correctly, and we therefore
began to investigate where the differences lie upon including
temperature effects.
2.3.2 The adiabatic case
We verified our adiabatic multi-layer models by comparing
them with those of Valencia et al. (2007), who constructed
similar models using the ice VII equation of state for water
(Fig. 5). When we set high surface pressures (1010 Pa) we
forced the surface layer of water to begin as ice VII or close
to it and therefore produced a very similar mass -- radius re-
lation. However, we predict inflated radii at lower surface
pressures and therefore conclude that surface temperature
and surface pressure are both important factors for deter-
mining the radius of a planet with a water layer. We further
explore this relationship in our results section.
There are minor differences between our mass -- radius re-
lations and the mass -- radius relations presented by Valencia
0246810Planet mass / M0.00.51.01.52.02.53.0Planet radius / RH2OMgSiO3FeSeager et al.This work10
Thomas & Madhusudhan
Figure 5. Validation of adiabatic models. Our mass -- radius re-
lations reproduce those for dry planets well, and predict inflated
radii for planets with water layers. Here we show mass -- radius re-
lations for two classes of models: dry planets (33 per cent Fe and
67 per cent MgSiO3 by mass), and wet planets (17 per cent Fe,
33 per cent MgSiO3, and 50 per cent water). We compared the
mass -- radius relations with the work of Valencia et al. (2007) who
constructed models with ice VII layers. At a surface pressure of
1010 Pa the water layer in the wet planets is mostly ice VII and
so our results are similar in this case. Small differences are likely
due to our different equation of state choice for ice VII. However,
at lower surface pressures, water can have an extended lower den-
sity shell that results in a larger planet than otherwise expected.
The surface temperature in these models is 550 K, matching the
characteristic temperature used by Valencia et al. (2007) in their
models.
et al. (2007). We slightly underpredict the radii of lower-
mass planets in models with surface pressures of 1010 Pa.
These differences are likely due to our choice of equation of
state: we use only simple isothermal prescriptions for iron
and magnesium silicate and include more phases of water
than just ice VII. We also did not include any treatment of
conductive boundary layers in our models. In general, how-
ever, our results agree well with theirs.
We also compared our results with the evolutionary
models of Lopez et al. (2012). Although we were able to
reproduce their mass -- radius relation for Earth-like planets,
we were less successful when adding extended water layers
(Fig. 6). We can match the radius of an arbitrary planet by
choosing an appropriate surface pressure but we underpre-
dict the radii of small planets and overpredict the radii of
large planets compared with their results. This may be a
result of different equation of state choices or different tem-
perature gradients during the course of their evolutionary
calculations.
Fig. 6 also provides a first indication of how changes in
surface temperature can affect the mass -- radius relation. We
highlight the magnitude of these differences and note that
they are still significant at pressures of 107 Pa (100 bar) and
up, well into the pressure region where many atmospheric
models terminate. We explore the effects on our models of
changing surface temperature, surface pressure and compo-
sition in the next section.
Figure 6. Comparison with evolutionary models. We plot dry
(Earth-like) and wet (50 per cent water on an Earth-ratio
core/mantle) mass -- radius relations. Shown for comparison are
models by Lopez et al. (2012), who build on work by Fortney
et al. (2007) and Nettelmann et al. (2011) by using a thermal
evolution approach to track the entropy within each planet as
it cools. Surface temperature significantly alters the mass -- radius
relation in our models. The surface temperature in these models
is 700 K but the shaded band shows models with surface temper-
atures from 500 to 900 K, a significant spread, which is caused by
temperature-dependent density changes of water at lower pres-
sures. We chose a surface pressure of 107 Pa to approximately
match the radii of Lopez et al. (2012). Their method does not
begin from an explicit surface pressure, as ours does.
3 RESULTS
We have explored the effects of temperature dependence on
the radii of water-rich super-Earths. This section shows that
significant radius variations can occur across temperature
ranges relevant to super-Earths. We explored the depen-
dence of super-Earth radii on three key model parameters.
1. Planet surface temperature, with the water layer tem-
perature profile taken as
a) isothermal, or
b) adiabatic.
2. Planet surface pressure.
3. Planet composition, i.e. water mass fraction.
3.1 Effect of surface temperature on isothermal
and adiabatic interiors
We found that thermal expansion can lead to significant
changes in the radii of water-rich super-Earths. We con-
structed super-Earths in two different ways. First we mod-
elled them as isothermal spheres containing an Earth-like
core (33 per cent Fe and 67 per cent MgSiO3) underneath a
water layer of 30 per cent of the planet's mass. Then we in-
stead allowed the temperature to increase adiabatically into
the water layer. Fig. 7 shows that the assumption that ther-
mal expansion effects are negligible, which was made in some
previous studies, is not the case. This is true in two senses.
First, a significant temperature dependence exists when we
adopt an adiabatic interior temperature profile compared
with an isothermal one. The surface temperature also af-
fects the radius of a planet within both types of models.
MNRAS 000, 1 -- 16 (2016)
0246810Planet mass / M0.00.51.01.52.02.5Planet radius / REarth-like50% H2O50% H2O107Pa1010PaValencia et al.This work0246810Planet mass / M0.00.51.01.52.02.53.0Planet radius / REarth-like50% H2O±200K surfaceLopez et al.This workThermal effects in super-Earth interiors
11
water layer may consist of supercritical fluid rather than
liquid, solid, or vapour (Fig. 1).
A significant dependence on surface temperature also
exists when using the adiabatic models. That is, changing
the surface temperature affects the radius of a model wa-
ter super-Earth even when its temperature profile is already
being treated as adiabatic. In the case of a 10 M⊕ planet,
increasing the surface temperature from 300 to 1000 K gave
a radius increase of 0.6 R⊕. For an Earth-mass planet the in-
crease was approximately 0.3 R⊕ for the same temperature
range.
We have highlighted above the change in the adiabatic
models, which we claim are a more realistic representation
of the actual temperature structure within the planet. But
even the isothermal models show a significant increase in ra-
dius with the planet's temperature. For a 10 M⊕ planet, the
change in radius is 0.3 R⊕ from 300 to 1000 K. This is due
to the thermal expansion of the planet as a whole, rather
than of one small part of the water layer near the surface.
We do not necessarily expect an adiabatic temperature gra-
dient throughout the whole planet because the entire inte-
rior may not all be convective. For example, Valencia et al.
(2007) included conductive boundary layers in their mod-
els. In that case, the true temperature-dependent behaviour
of the mass -- radius diagram might lie between the adiabatic
and isothermal cases. Despite this, Fig. 7 shows that the
surface temperature can still play an important role in de-
termining the radius of a planet if it has a substantial water
layer. This is true even in the extreme isothermal case where
there is no temperature gradient at all within the planet.
These models have a surface pressure of 107 Pa (100 bar)
so this effect is not due to the strong liquid -- vapour transi-
tion at 1 bar. In fact, we still see these effects past the critical
pressure of water (2.206×107 Pa). The critical point, which is
visible in Figs 1 and 2, is the point in temperature -- pressure
space beyond which there is no distinct phase transition from
liquid to vapour. This indicates that a liquid -- vapour transi-
tion is not required to produce a significantly inflated radius
when the water layer is heated. We discuss the effect of pres-
sure on these models further in the next section.
3.2 Effect of surface pressure
The surface pressure can strongly affect the temperature-
dependent thickness of the water layer (Fig. 8). For example,
at high temperatures (1000 K), increasing the surface pres-
sure of a 10 per cent water and 4 M⊕ planet from 10 bar to
1000 bar compresses the water layer significantly, decreasing
the planet's radius by a factor of two. And at low pressures
we see a bifurcation in the surface pressure contours where a
surface temperature increase of 100 K or less can inflate the
radius of a watery super-Earth by more than 50 per cent.
This is the result of a transition across the liquid -- vapour
phase boundary, which exists at pressures up to the criti-
cal pressure of water (2.206× 107 Pa). Our interior structure
code is most likely not the best choice for modelling such a
quasi-atmospheric layer: we did not handle radiative energy
transfer in our models. We have therefore not undertaken a
detailed study of the behaviour of these vapour layers. They
likely require a more sophisticated treatment of the temper-
ature profile than our adiabatic assumption.
Despite observing highly inflated radii when the tem-
Figure 7. Dependence of watery super-Earth radii on surface
temperature and internal temperature profile. An increased sur-
face temperature results in an increased planetary radius. This
effect is especially pronounced in the full adiabatic temperature
treatment. Here we show super-Earths with an Earth-like core
under a 30 per cent water layer by mass. We treated the temper-
ature in two different ways: an isothermal treatment with a fixed
constant temperature and an adiabatic treatment where we fixed
the surface temperature but allowed the temperature to increase
inwards according to the adiabatic relation (equation 6). The adi-
abatic models are warmer and therefore significantly larger over-
all, but even the isothermal planets display some radius change
due to temperature. The effects of this temperature dependence
are comparable to current uncertainties on measured masses and
radii for some of the best-characterised exoplanets11. The surface
pressure in these models is 107 Pa (100 bar), and the temperature
increases in steps of 100 K. The large gap between 500 and 600 K
in the adiabatic case is due to a density discontinuity between the
liquid and vapour phases.
The adiabatic models have a larger radius for a given
mass when compared with the isothermal case. This is to be
expected: the average temperature is higher along an adia-
bat than an isotherm fixed at the surface temperature, and
the density of water generally decreases with temperature.
The increase in radius is significant at higher surface tem-
peratures, as shown in Fig. 7. For example, a 4 M⊕ 30 per
cent water planet with a 600 K surface has a radius of 1.8 R⊕
if its water layer is isothermal, but 2 R⊕ if it is adiabatic.
Across the super-Earth mass range we considered, the adi-
abatic radii increased by up to 0.3 R⊕ when compared with
the isothermal case. The difference becomes particularly pro-
nounced at higher surface temperatures, at which point the
11 These data are from exoplanets.org. We selected planets with
known radii and masses of 1 to 10 M⊕. We then plotted the twelve
planets with the lowest summed relative uncertainty in mass and
radius (∆R/R + ∆M/M).
MNRAS 000, 1 -- 16 (2016)
0246810Planet mass / M1.01.52.02.53.0Planet radius / RAdiabatic1000 K300 KIsothermal300 K1000 KBest measuredsuper-Earths12
Thomas & Madhusudhan
Figure 8. Dependence of radii on surface pressure. The effect of temperature on the radius of watery planets decreases with increasing
surface pressure, but remains significant (greater than about 0.1 R⊕) for pressures below 1000 bar. Here we show mass -- radius relations
for spheres with an Earth-like core under a 30 per cent water layer, changing only the surface pressure each time. The temperature
dependence remains even beyond the critical pressure of water (2.206× 107 Pa), at which point the surface water exists as a supercritical
fluid. Only at very high pressures (109 or 1010 Pa; 10000 or 100000 bar) does this temperature dependence vanish.
perature is increased across the liquid -- vapour phase bound-
ary, we still see temperature-dependent variation in the
planet's radius past the critical pressure of water. This is
because the density of water is still strongly temperature-
dependent in the super-critical regime. In fact, we might
reasonably expect the same inflated radii in any situation
where the pressure of the water layer places it in a region
of the water phase diagram that has significant tempera-
ture dependence. If the water layer is heated to thousands
of Kelvin, this temperature dependence may only begin to
disappear around 1010 Pa (100000 bar, Fig. 3). At a pressure
of 108 Pa (1000 bar), a watery super-Earth with a surface
temperature of 1000 K still has a radius that is up to 0.1 R⊕
larger than one with a surface temperature of 300 K. This is
comparable to or greater than the best current uncertainties
on measured super-Earth radii (Fig. 7), and indicates that
the surface temperature is a key parameter to consider when
one attempts to model planets with significant water mass.
We included no atmospheric layers in these models.
Other studies have provided more complete treatments of
atmospheric layers. For example, Rogers & Seager (2010b)
included a gas layer on top of an isothermal interior struc-
ture model in order to interpret the structure of the planet
GJ 1214b. And Valencia et al. (2013) used internal struc-
ture models coupled with an atmospheric layer, exploring
the dependence of radii on various model parameters in-
cluding equilibrium temperature and water content. Given
that we set the surface pressure to between 105 and 1010 Pa
(1 and 100000 bar), our models must therefore represent the
layers interior to an atmosphere of some sort.
3.3 Effect of water content
We find that changing the water content does not sig-
nificantly affect the temperature-dependent behaviour dis-
cussed in earlier sections (Fig. 9). We constructed planets
with water, silicate, and iron layers, fixing the silicate:iron
mass ratio to the Earth value of 2:1 and allowing the water
shell to vary in mass. These models correspond to an Earth-
like nucleus with an extended water layer at the surface.
The effects of surface temperature on radius are compa-
rable in magnitude across all our models with water layers,
MNRAS 000, 1 -- 16 (2016)
123456Surface pressure 1 bar300 K surface1000 K surface10 bar0246810123456100 bar2468101000 barPlanet mass / MPlanet radius / Reven when we set the water layer mass to just 1 per cent of
the mass of the entire planet. For a 10 M⊕ super-Earth with a
surface pressure of 107 Pa (100 bar), the radial change when
the surface temperature increases 300 to 1000 K is 0.5 R⊕
(for a 50 per cent water planet) and 0.4 R⊕ (for a 1 per cent
water planet). This similarity holds across the entire range of
planetary masses we considered. This suggests that the bulk
of the radius change comes from a water layer on the surface
whose density depends strongly on the surface temperature.
3.4 Effect of temperature dependence on phase
structure
for water
Our adiabatic assumption provides
layers
which span different phases depending on the pressure --
temperature profile. As previously noted, we observed a sig-
nificant bifurcation in the mass -- radius diagrams when the
surface temperature crossed the condensation curve of wa-
ter. As an example of how the surface temperature affects
the structure of a planet at pressures beyond the critical
pressure of water, Fig. 10 shows pressure -- temperature pro-
files for adiabatic spheres of water with different surface tem-
peratures. The planet contains a significant ice VII com-
ponent when the surface temperature is low. But at high
surface temperatures the centre of the planet may consist
mostly of the superionic or plasma phases of water, which
are shown in Figs 1 and 2.
Others have explored the layered phase structure of wa-
tery planets (e.g. Zeng & Sasselov 2014; Ehrenreich et al.
2006). In particular, Ehrenreich et al. (2006) included an
analysis of radiogenic heating in their models to assess the
feasibility of having a liquid ocean under a cold ice shell.
Though we have not assessed how the layered phase
structure of a planet's water layer might affect other prop-
erties of the planet, the phase of water could be important
for its potential to sustain convective energy transport or
magnetic fields (Zeng & Sasselov 2014). We did include a
more complete treatment of the thermal expansion coeffi-
cient α (calculating it directly from the equation of state)
and the variable heat capacity cp. This approach may result
in a phase structure that differs from other studies. In fu-
ture we anticipate investigating this more closely to assess
whether these internal energy sources are indeed sufficient to
drive convection throughout our models: is the assumption
of a fully convective interior reasonable? As a first indica-
tion of this, we consider the fact that we do not find major
deviations from the mass -- radius relations of Valencia et al.
(2007) (Fig. 5) to be promising. This is despite the fact that
they include conductive boundary layers in their models.
The phase structure is also of interest when we consider
questions of habitability. The properties of water change
significantly near the critical point: water becomes a low-
dielectric fluid and a poor solvent for polar substances (An-
simov et al. 2004). The nature of reactions supported by wa-
ter is also expected to change at high temperatures (Kruse
& Dinjus 2007). We would expect significant changes in any
kind of life that could be found within these water layers,
both when compared with liquid oceans on Earth and when
seen over time as the planet's structure evolved.
MNRAS 000, 1 -- 16 (2016)
Thermal effects in super-Earth interiors
13
4 CONCLUSION AND DISCUSSION
In this paper we have presented planetary interior structure
models of water-rich super-Earths. The models incorporate
a temperature-dependent water equation of state and use
an adiabatic treatment for the temperature gradient. In do-
ing so, we synthesized an updated equation of state for wa-
ter which attempts to capture all the relevant temperature-
dependent behaviour. We directly calculated the thermal ex-
pansion coefficient α from the equation of state, rather than
treating it as a constant, and we used a variable heat ca-
pacity based on experimental data. Our conclusions are as
follows.
First, when one models a solid planet, adding a wa-
ter layer comes with a substantial thermal dependence. By
this we mean that the temperature of the planet may sub-
stantially alter the radius of the planet as the water layer
expands and contracts. Previous studies have shown that
including a temperature gradient in Earth-like planets pro-
duces a minimal change in its radius (Howe et al. 2014; Gras-
set et al. 2009; Seager et al. 2007). We showed that this
assumption no longer holds once large water layers are con-
sidered, even setting aside the unrealistic case of a 100 per
cent water planet. For example, consider the case of a 4 M⊕
planet with an Earth-like core underneath a water layer of 5
per cent of the planet's total mass. If the surface pressure is
107 Pa (100 bar), the difference in the planet's radius when
the surface is heated from 300 K to 1000 K is approximately
0.3 R⊕ (Fig. 9). This effect is on top of any thermal expan-
sion of iron and silicate: our models treated the rocky layers
as isothermal. It is also in addition to any uncertainty in the
equation of state itself. Such changes in radii are significant
considering that current observations can already measure
super-Earth radii to precisions better than 0.1 R⊕ (e.g. Fig.
7).
The strength of the planet radius-temperature relation
also depends on the surface pressure. This is a result of the
decreasing thermal expansion of water with pressure: the
coefficient of thermal expansion is much smaller in high-
pressure ice than in the liquid, vapour, or supercritical fluid
phases. At pressures of more than about 1010 Pa (100000 bar)
the radial temperature dependence becomes irrelevant: the
uncertainty in current planetary radius measurements is
larger than any conceivable radial change owing to tem-
perature effects, so more precise structural models may not
be useful. However, there is still a significant radial depen-
dence on temperature at lower surface pressures. At 108 Pa
(1000 bar), a watery super-Earth with a surface temperature
of 1000 K can be up to 0.1 R⊕ larger than one with a surface
temperature of 300 K. It is therefore important to include
temperature effects in the interior models if an accurate ra-
dius is required as part of the model.
This pressure dependence manifests itself most strongly
below the critical point of water. At pressures below this crit-
ical pressure, a phase transition still exists between liquid
and vapour. There is therefore a bifurcation in the mass --
radius diagram: a small increase in surface temperature can
causes a large change in radius (up to a factor of two) as
the surface water vaporises. We caution that it is likely not
appropriate to attempt to treat such vapour layers using our
approach, which is intended for interior structures. However,
14
Thomas & Madhusudhan
Figure 9. Dependence of radii on water mass fraction. Even low-mass water layers result in planets that are strongly affected by
temperature changes, especially when water on the surface is hot enough to be in the vapour or supercritical phase. Here we show mass --
radius relations for multi-layer planets: an iron core with silicate and (in all but the first panel) water layers. We show the Earth-like
iron-silicate core in each panel for comparison. All the watery planets are larger than the dry case owing to the lower density of water.
Surface temperature variation affects the radius of a watery planet by a similar amount in each case, and it can increase the radius by
up to 25 per cent. Because the iron and silicate layers are isothermal, this variation is due solely to temperature effects in the water
layer. We fixed the silicate:iron mass ratio at 2:1 and set the surface pressure to 107 Pa (100 bar). The temperature contours are in steps
of 100 K.
a lesser version of this effect is still visible at higher pres-
sures.
We consider the surface pressure as a free parameter
in our models. In principle, the surface pressure could be
constrained through spectroscopic observations of the plan-
etary atmosphere, though such observations are currently
difficult for super-Earths. The surface pressure is set by
the depth beyond which atmospheric measurements can no
longer probe. Madhusudhan & Redfield (2015) discussed
planets with water-rich atmospheres, describing the use of
measurements both in and out of opacity windows to deter-
mine the atmospheric thickness. The pressure to which these
measurements probe varies from 0.1 bar (in regions of high
opacity; that is, outside an atmospheric window) to 100 bar
(within such a window). Our models go beyond this pressure
range, and are therefore appropriate to treat the structure
of the planet below the observable opacity surface.
In the case of a volatile layer such as water, the line
between interior and atmosphere can become blurred. The
picture is complicated by atmospheric effects that can in-
crease the opacity. If a cloud layer forms in the atmosphere,
the opacity surface may not necessarily be at the same depth
or pressure as any solid surface of the planet. Turbidity ef-
fects around the critical point may also affect the opacity. It
is for this reason that high-temperature exoplanets are in-
teresting: at higher temperatures, a cloud deck is less likely
to occur and atmospheric measurements are therefore able
to probe deeper. The previously-mentioned opacity windows
may therefore be able to provide a view through the atmo-
sphere to the planet's surface, or at least to a point where
the assumption of interior convective mixing is more likely
to hold.
We therefore conclude that, in some cases, planetary
heating may alter the interpretation of a planet's radius if
a water layer is part of the model. This is especially true if
the planet consists entirely of water, but this is an unlikely
physical scenario. More importantly, the result is still signif-
icant even if the surface of the water layer is at moderately
MNRAS 000, 1 -- 16 (2016)
0.51.01.52.02.53.0Earth-like (isothermal 300 K)300 K surface1000 K surface1% H2O5% H2O02468100.00.51.01.52.02.53.010% H2O24681030% H2O24681050% H2OPlanet mass / MPlanet radius / RThermal effects in super-Earth interiors
15
mate into this approach should therefore give better con-
straints.
From an observational perspective, these results are
most interesting at intermediate pressures. At low pressures
(105 Pa or 1 bar) we cannot claim that we accurately capture
the behaviour of what is now essentially an atmosphere, be-
cause we include no prescription for radiative energy trans-
port in our models. At high pressures (1010 Pa or 100000 bar)
any temperature dependence in the water equation of state
disappears. The physical scenario most relevant for these
models is therefore that of a water layer (ocean, ice or su-
percritical fluid) underneath a thin or moderate atmosphere.
Others such as Rogers & Seager (2010b) have already in-
cluded volatile layers on top of interior structure models.
Adding more complete temperature dependence to the inte-
rior portion of these planetary models is a worthwhile future
direction if we wish to treat them as water-rich.
We look forward to two developments in particular.
The first is improved atmospheric characterisation and mod-
elling, which will provide useful pressure and temperature
boundary conditions at the base of the atmosphere. The
question of interior -- atmospheric interactions is a rich one
that is only starting to be explored. Integrating atmospheric
and interior models promises progress on questions about
surface chemistry, outgassing and other processes that can
shape the atmosphere of a planet. The second development
that will make use of this work is improved spectral resolu-
tion of atmospheric observations, and in particular the abil-
ity to seek out atmospheric windows (Madhusudhan & Red-
field 2015). By observing at wavelengths which pass through
the atmosphere, we can in principle directly measure the ra-
dius of any solid interior underneath that atmosphere and
thus have a better starting point for interpreting the interior
structure.
ACKNOWLEDGEMENTS
We thank the anonymous reviewer for an insightful review
and Christopher Tout for helpful discussions and comments.
ST gratefully acknowledges support from the Royal Society
of New Zealand.
REFERENCES
Ansimov M. A., Sengers J. V., Sengers J. M. L., 2004,
in
Fernandez-Prini R., Harvey A., Palmer D., eds, , Aqueous
Syst Elev Temp Press. Elsevier Academic Press, London,
Chapt. 2, p. 29
Baraffe I., Chabrier G., Barman T. S., 2008, A&A, 482, 315
Baraffe I., Chabrier G., Fortney J. J., Sotin C., 2014, in Beuther
H., Klessen R. S., Dullemond C. P., Henning T., eds, , Pro-
tostars Planets VI. University of Arizona Press, Tucson, pp
763 -- 786, doi:10.2458/azu uapress 9780816531240
Belonoshko A., Saxena S., 1991, Geochim Cosmochim Acta, 55,
381
Broeg C., et al., 2013, EPJ Web Conf, 47, 03005
Choukroun M., Grasset O., 2007, J Chem Phys, 127, 124506
Dorn C., Khan A., Heng K., Connolly J. A. D., Alibert Y., Benz
W., Tackley P., 2015, A&A, 577, A83
Figure 10. Model pressure -- temperature profiles. Increasing the
surface temperature means that more of the planet consists of
superionic or plasma phases of water, with the transition to high-
pressure ice happening deeper within the interior or not at all.
Here we show pressure -- temperature profiles for 3 M⊕ spheres of
water with a surface pressure of 5× 107 Pa, which is beyond the
critical pressure. At a surface temperature of 300 K, the planet
consists of liquid water over an ice VII core. But the ice VII
phase may not be present within the interior at higher surface
temperatures. Instead, the bulk of the planet consists of water
in the superionic or plasma state. This, combined with the low-
density supercritical fluid at the surface, results in an inflated
radius.
high pressures and lies underneath a heavy atmosphere. All
that is required for the water layer's density to change signif-
icantly from the isothermal case is for a temperature increase
of a few hundred Kelvin. Moreover, even isothermal watery
planets have some degree of radial temperature dependence:
up to 0.3 R⊕ across the mass range of super-Earths and in
the temperature range of 300 to 1000 K.
Understanding how the mass -- radius relation can be af-
fected by temperature allows us to take the step of detecting
and characterising water-rich planets, taking their surface
temperatures into account while modelling them. This is an
important precursor to narrow the search to planets that
would be considered more classically habitable. It will be
especially useful in the context of the next generation of
super-Earths expected to be found orbiting bright stars by
missions such as PLATO (Rauer et al. 2014), TESS (Ricker
et al. 2014) and CHEOPS (Broeg et al. 2013). This approach
is promising because it is linked to the characteristic equi-
librium temperature, which can be determined from obser-
vations of the planet, and so can be included in analyses of
populations of planets. Through this we might better under-
stand what proportion of planets include substantial water
content.
The temperature dependence is also important to take
into account in approaches such as that of Kipping et al.
(2013), where a watery interior model is used to place a lower
bound on the atmospheric height of an observed planet. We
have shown that the radius of an adiabatic watery planet
may be significantly higher than the zero-temperature or
isothermal case. Incorporating a surface temperature esti-
MNRAS 000, 1 -- 16 (2016)
10710810910101011Water pressure / Pa102103Water temperature / KIhVVIIX300 K500 K800 K1000 K16
Thomas & Madhusudhan
Dunaeva A. N., Antsyshkin D. V., Kuskov O. L., 2010, Sol Syst
Sugimura E., Komabayashi T., Hirose K., Sata N., Ohishi Y.,
Res, 44, 202
Ehrenreich D., Lecavelier des Etangs A., Beaulieu J., Grasset O.,
2006, ApJ, 651, 535
Fei Y., Mao H.-k., Hemley R. J., 1993, J Chem Phys, 99, 5369
Feistel R., Wagner W., 2006, J Phys Chem Ref Data, 35, 1021
Fortney J. J., Nettelmann N., 2009, Space Sci Rev, 152, 423
Fortney J. J., Marley M. S., Barnes J. W., 2007, ApJ, 659, 1661
Frank M. R., Fei Y., Hu J., 2004, Geochim Cosmochim Acta, 68,
2781
French M., Mattsson T., Nettelmann N., Redmer R., 2009, PRB,
79, 054107
Grasset O., Schneider J., Sotin C., 2009, ApJ, 693, 722
Guillot T., 1999, Planet Space Sci, 47, 1183
Haghighipour N., 2011, Contemp Phys, 52, 403
Howe A. R., Burrows A. S., Verne W., 2014, ApJ, 787, 26
Hubbard W. B., MacFarlane J. J., 1980, J Geophys Res, 85, 225
Hubbard W., Marley M. S., 1989, Icarus, 78, 102
Kipping D. M., Spiegel D. S., Sasselov D. D., 2013, MNRAS, 434,
1883
Knudson M. D., Desjarlais M. P., Lemke R. W., Mattsson T. R.,
French M., Nettelmann N., Redmer R., 2012, PRL, 108,
091102
Kruse A., Dinjus E., 2007, J Supercrit Fluids, 39, 362
Levi A., Sasselov D., Podolak M., 2014, ApJ, 792, 125
Lopez E. D., Fortney J. J., Miller N., 2012, ApJ, 761, 59
Lyon S. P., Johnson J. D., 1992, Technical report, SESAME: The
Los Alamos National Laboratory equation of state database.
LA-UR-92 3407, Los Alamos National Lab
Madhusudhan N., 2012, ApJ, 758, 36
Madhusudhan N., Redfield S., 2015, Int J Astrobiol, 14, 177
Madhusudhan N., Lee K. K. M., Mousis O., 2012, ApJ, 759, L40
Milone E. F., Wilson W. J., 2014, Solar System Astrophysics:
Background Science and the Inner Solar System, 2 edn.
Springer, New York, doi:10.1007/978-1-4614-8848-4
Mordasini C., Alibert Y., Georgy C., Dittkrist K.-M., Klahr H.,
Henning T., 2012, A&A, 547, A112
More R. M., Warren K. H., Young D. A., Zimmerman G. B., 1988,
Phys Fluids, 31, 3059
Nettelmann N., Holst B., Kietzmann A., French M., Redmer R.,
Blaschke D., 2008, ApJ, 683, 1217
Nettelmann N., Fortney J. J., Kramm U., Redmer R., 2011, ApJ,
733, 2
Owen J. E., Wu Y., 2015, preprint (arXiv:1506.02049)
Poirier J.-P., 2000, Introduction to the Physics of the Earth's
Interior, 2nd edn. Cambridge University Press, Cambridge
Press W. H., 2007, Numerical Recipes: The Art of Scientific Com-
puting, 3rd edn. Cambridge University Press, Cambridge
Rauer H., et al., 2014, Exp Astron, 38, 249
Redmer R., Mattsson T. R., Nettelmann N., French M., 2011,
Icarus, 211, 798
Ricker G. R., et al., 2014, in Oschmann J. M., Clampin M., Fazio
G. G., MacEwen H. A., eds, , Vol. 9143, Proc SPIE. pp
914315 -- 914320 (arXiv:1406.0151), doi:10.1117/12.2063489,
http://arxiv.org/abs/1406.0151
Rogers L. A., 2015, ApJ, 801, 41
Rogers L. A., Seager S., 2010a, ApJ, 712, 974
Rogers L. A., Seager S., 2010b, ApJ, 716, 1208
Rogers L. A., Seager S., 2012, Phd thesis, MIT
Salpeter E., Zapolsky H., 1967, PR, 158, 876
Seager S., Kuchner M., Hier-Majumder C. A., Militzer B., 2007,
ApJ, 669, 1279
Senft L. E., Stewart S. T., 2008, Meteorit Planet Sci, 43, 1993
Sotin C., Grasset O., Mocquet A., 2007, Icarus, 191, 337
Sotin C., Jackson J. M., Seager S., 2010, in Seager S., ed., , Exo-
planets. University of Arizona Press, Chapt. 16, pp 375 -- 395
Stewart S. T., Ahrens T. J., 2005, J Geophys Res, 110, E03005
Dubrovinsky L. S., 2010, PRB, 82, 134103
Thompson S. L., Lauson H. S., 1972, Technical report, Improve-
ments in the Chart B radiation hydrodynamic code III: re-
vised analytic equations of state. SC-RR-71 0714, Sandia Na-
tional Lab
Unterborn C. T., Dismukes E. E., Panero W. R., 2015, preprint
(arXiv:1510.07582)
Valencia D., O'Connell R. J., Sasselov D. D., 2006, Icarus, 181,
545
Valencia D., Sasselov D. D., O'Connell R. J., 2007, ApJ, 665, 1413
Valencia D., Ikoma M., Guillot T., Nettelmann N., 2010, A&A,
516, A20
Valencia D., Guillot T., Parmentier V., Freedman R. S., 2013,
ApJ, 775, 10
Vazan A., Kovetz A., Podolak M., Helled R., 2013, MNRAS, 434,
3283
Vinet P., Smith J., Ferrante J., Rose J., 1987, PRB, 35, 1945
Wagner W., Pruss A., 2002, J Phys Chem Ref Data, 31, 387
Wagner F., Sohl F., Hussmann H., Grott M., Rauer H., 2011,
Icarus, 214, 366
Wilson H. F., Militzer B., 2012, ApJ, 745, 54
Wilson H., Wong M., Militzer B., 2013, PRL, 110, 151102
Zapolsky H. S., Salpeter E. E., 1969, ApJ, 158, 809
Zeng L., Sasselov D. D., 2013, PASP, 125, 227
Zeng L., Sasselov D. D., 2014, ApJ, 784, 96
This paper has been typeset from a TEX/LATEX file prepared by
the author.
MNRAS 000, 1 -- 16 (2016)
|
1908.05784 | 2 | 1908 | 2019-08-20T05:12:58 | Dusty clumps in circumbinary discs | [
"astro-ph.EP"
] | Recent observations have revealed that protoplanetary discs often exhibit cavities and azimuthal asymmetries such as dust traps and clumps. The presence of a stellar binary system in the inner disc regions has been proposed to explain the formation of these structures. Here, we study the dust and gas dynamics in circumbinary discs around eccentric and inclined binaries. This is done through two-fluid simulations of circumbinary discs, considering different values of the binary eccentricity and inclination. We find that two kinds of dust structures can form in the disc: a single horseshoe-shaped clump, on top of a similar gaseous over-density; or numerous clumps, distributed along the inner disc rim. The latter features form through the complex interplay between the dust particles and the gaseous spirals caused by the binary. All these clumps survive between one and several tens of orbital periods at the feature location. We show that their evolution strongly depends on the gas-dust coupling and the binary parameters. Interestingly, these asymmetric features could in principle be used to infer or constrain the orbital parameters of a stellar companion - potentially unseen - inside the inner disc cavity. Finally, we apply our findings to the disc around AB Aurigae. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- ?? (2019)
Preprint 21 August 2019
Compiled using MNRAS LATEX style file v3.0
Dusty clumps in circumbinary discs
Pedro P. Poblete1,2, Nicolás Cuello1,2 and Jorge Cuadra1,2
1Instituto de Astrofísica, Pontificia Universidad Católica de Chile, Santiago, Chile,
2Núcleo Milenio de Formación Planetaria (NPF), Chile.
Accepted 2019 August 15. Received 2019 August 14; in original form 2019 April 25
ABSTRACT
Recent observations have revealed that protoplanetary discs often exhibit cavities and az-
imuthal asymmetries such as dust traps and clumps. The presence of a stellar binary system
in the inner disc regions has been proposed to explain the formation of these structures. Here,
we study the dust and gas dynamics in circumbinary discs around eccentric and inclined bina-
ries. This is done through two-fluid simulations of circumbinary discs, considering different
values of the binary eccentricity and inclination. We find that two kinds of dust structures can
form in the disc: a single horseshoe-shaped clump, on top of a similar gaseous over-density;
or numerous clumps, distributed along the inner disc rim. The latter features form through the
complex interplay between the dust particles and the gaseous spirals caused by the binary. All
these clumps survive between one and several tens of orbital periods at the feature location.
We show that their evolution strongly depends on the gas -- dust coupling and the binary param-
eters. Interestingly, these asymmetric features could in principle be used to infer or constrain
the orbital parameters of a stellar companion -- potentially unseen -- inside the inner disc
cavity. Finally, we apply our findings to the disc around AB Aurigae.
Key words: protoplanetary discs -- planets and satellites : formation -- hydrodynamics --
methods: numerical.
1
INTRODUCTION
In the last years, the field of planet formation has experienced an
unprecedented development thanks to last-generation telescopes. In
particular, by combining multi-wavelength observations of proto-
planetary discs, it has been possible to map the dust distribution
for a wide range of grain sizes around young stars. Extreme adap-
tive optics instruments in large optical/NIR telescopes and radio
antennas observing at mm and submm wavelengths played a key
role in achieving this task. This shed some light on the very first
stages of planet formation within these systems (Avenhaus et al.
2018; Pinilla et al. 2018). Among the now overwhelming number
of ALMA observations of protoplanetary discs, the continuum emis-
sion detected around HL Tau is one of the most spectacular (ALMA
Partnership et al. 2015). Especially, the numerous gaps observed
suggest that planets might have already formed in this young disc
(Dipierro et al. 2015b). Besides the routinely detected gaps, there
are also numerous observations of enigmatic structures such as
rings, spirals, warps, clumps, and vortices. For instance, the recent
DSHARP survey by Andrews et al. (2018) mapped twenty nearby
protoplanetary discs at an astonishing resolution of roughly 5 au.
However, the rich structure of these systems remains only partly
understood from the theoretical point of view (Armitage 2018).
Interestingly, a fraction of these circumstellar discs orbit a bi-
nary stellar system instead of a single star. These constitute a spe-
cial category of protoplanetary discs called circumbinary. Consid-
ering the stellar context, roughly half of the solar-type stars are
© 2019 The Authors
singles, whereas about 33% of them form double systems (Ragha-
van et al. 2010; Tokovinin 2014). Therefore, about a third of the
young stellar systems could potentially harbour circumbinary discs
(CBDs), along with circumstellar ones. Hence, a proper under-
standing of circumbinary disc dynamics is of crucial importance
(Nixon et al. 2013; Dunhill et al. 2015).
The case of the disc around HD 142527 is particularly enlight-
ening in this regard. Fukagawa et al. (2006) first detected a disc
with several spiral arms and a large inner cavity of roughly 90 au.
This disc was initially thought to be orbiting a single star. However,
a companion was later discovered inside the inner cavity by Biller
et al. (2012). The stellar masses in HD 142527 are 1.8 M(cid:12) (Gaia
Collaboration et al. 2016) and 0.4 M(cid:12) (Christiaens et al. 2018), so
it is an unequal-mass binary. Further studies focused on the com-
panion's orbital motion (Lacour et al. 2016; Claudi et al. 2019) in-
dicating that the binary is eccentric and likely inclined with respect
with the disc. Based on these constraints, Price et al. (2018b) pre-
sented a consistent hydrodynamical model of HD 142527 where
the CBD is periodically perturbed by the inner binary. Remarkably,
the resulting gaseous and dust structures are in excellent agreement
with all the available multi-wavelength observations: i) the spirals
and their location (Avenhaus et al. 2014; Christiaens et al. 2014),
ii) the cavity size (Perez et al. 2015), iii) the dusty clumps along
a horseshoe (Casassus et al. 2015b; Boehler et al. 2017), iv) the
gaseous filaments crossing the cavity (Casassus et al. 2013), and v)
the shadows (Avenhaus et al. 2014) -- likely caused by the pres-
ence of a misaligned inner disc (Marino et al. 2015).
2
P.P. Poblete, N. Cuello & J. Cuadra
This circumbinary scenario could very well apply to other
discs with large inner cavities exhibiting various asymmetries. For
instance, Ragusa et al. (2017) explored how unequal-mass (circu-
lar) binaries in a coplanar configuration are able to generate lop-
sided features and horseshoes at the edge of the cavity -- compara-
ble to the ones observed. Alternatively, the presence of vortices has
been widely proposed to explain the same asymmetries in proto-
planetary discs (Meheut et al. 2012; Lyra & Lin 2013; Ataiee et al.
2013; van der Marel et al. 2016). It is worth noting that, in the bi-
nary scenario, no vortex is required whatsoever. Regardless of their
origin, these azimuthal pressure maxima are expected to efficiently
trap dust in the inner disc regions (Birnstiel et al. 2013).
The aim of this work is to study the effect of an inclined and
eccentric inner binary on the dust content of the surrounding CBD.
To do so, we consider relatively high eccentricities (eB = 0.5 and
eB = 0.75) and different inclinations for the binary, from prograde
(iB = 0°) to retrograde (iB = 180°) configurations. The numerical
method and the initial setup of our three-dimensional hydrodynam-
ical simulations are described in Section 2. We report our results in
Section 3. In Section 4, we discuss the formation of dusty clumps,
their evolution and how these can be used to infer the presence of a
potentially unseen inner companion. Finally, we draw our conclu-
sions in Section 5.
2 NUMERICAL METHOD
We perform 3D hydrodynamics simulations of circumbinary discs
(CBDs) using the PHANTOM smoothed particle hydrodynamics
(SPH) code (Price et al. 2018a). We use the two-fluid method
in order to model the interaction between gas and dust particles
as described in Laibe & Price (2012a,b). Although each fluid is
treated independently, gas and dust particles interact with each
other through aerodynamical drag forces. This means that the back-
reaction from the dust on the gas is included in our calculations.
2.1 Binary setup
We consider a binary system where both stars are treated as sink
particles (Bate et al. 1995). We explore a range of orbital pa-
rameters similar to those observed in HD 142527. In particular,
we test different combinations of binary inclination (iB) and ec-
centricity (eB). The mass ratio between the primary and the sec-
ondary stars is fixed at q = M1/M2 = 0.25, with M1 =
2 M(cid:12) and M2 = 0.5 M(cid:12). The semi-major axis is set to 40 au
(as in Price et al. (2018b) for HD 142527B). The free param-
eters in our simulations are eB and iB. In this work, we con-
sider the following sets of values: eB = {0.50, 0.75} and iB =
{0◦, 30◦, 60◦, 90◦, 120◦, 150◦, 180◦}. We model each system for
a hundred binary orbits. During this evolutionary time the binary
orbit does not change significantly, as expected. Therefore, the pa-
rameters aforementioned can be considered as constant throughout
each simulation. The effect of the different combinations of orbital
parameters on the CBD structure will be discussed in more detail
in Section 2.3.
2.2 Disc setup
2.2.1
Initial conditions
Figure 1. Parameter space of orbits modelled. The left panel shows face-on
views of the two simulated eccentricities, eB = 0.50 in red and eB = 0.75
in blue. A circular orbit is shown with a dotted line for comparison. The
right panel shows the projection of the orbit with eB = 0.75 for all eight
modelled inclinations.
Orbit name
e50-i0
e50-i30
e50-i60
e50-i90
e50-i120
e50-i150
e50-i180
e75-i0
e75-i30
e75-i60
e75-i90
e75-i120
e75-i150
e75-i180
eB
0.50
0.50
0.50
0.50
0.50
0.50
0.50
0.75
0.75
0.75
0.75
0.75
0.75
0.75
iB
0°
30°
60°
90°
120°
150°
180°
0°
30°
60°
90°
120°
150°
180°
Table 1. List of orbital parameters for each simulation. The parameters are
the orbit name, eccentricity eB, and inclination with respect to the disc iB.
disc with 106 gas particles and 105 dust particles, assuming a total
gas mass of 0.01 M(cid:12) and a dust-to-gas ratio of 0.01. The spatial
distribution of both fluids is initially the same. We set the inner and
the outer edges at Rin = 90 au and Rout = 350 au, respectively.
The surface density profile is given by a power law, Σ ∝ R−1. The
temperature profile follows a shallower power law, T ∝ R−0.3, as
suggested by Casassus et al. (2015a) for this system. This corre-
sponds to H/R = 0.06 at Rin and H/R = 0.1 at Rout. We set the
SPH viscosity parameter αAV ≈ 0.3, which gives a mean Shakura --
Sunyaev disc viscosity of αSS ≈ 0.005 (Shakura & Sunyaev 1973)
(see Sect. C). We do not consider self-gravity effects nor magnetic
fields.
2.2.2 Dust modelling
The dust coupling to the gas is well described by the dimensionless
quantity called Stokes number (noted St). This quantity is defined
as the ratio of the orbital time-scale to the drag stopping time1. In
particular, when St (cid:28) 1 (St (cid:29) 1) then dust particles are strongly
We follow Price et al. (2018b) and consider a disc setup consistent
with the observations of HD 142527. We model the circumbinary
1 the time-scale for the drag to damp the local differential velocity between
the gas and dust.
MNRAS 000, 1 -- ?? (2019)
(weakly) coupled to the gas. Interestingly, if St ∼ 1, then the par-
ticles are marginally coupled to the gas and experience the fastest
radial drift (Weidenschilling 1977; Nakagawa et al. 1986). In all
our simulations, the mean free path of the dust particles is greater
than their size. Hence, the drag force falls in the Epstein regime
(Epstein 1924) and the Stokes number is given by
(cid:114) πγ
8
St =
ρss
ρcsf
Ωk,
(1)
where Ωk is the Keplerian angular velocity, ρs is the dust grain
density 2, cs the sound speed, s the grain size, f a correction factor
for supersonic drag, and ρ the total density ρg + ρd (i.e. the sum of
the gas and dust volume densities). This prescription is similar to
the treatment made by Dipierro et al. (2015b, 2016) and Price et al.
(2018b). Here, for computational convenience, we drop the ρd term
in the sum and use instead ρ = ρg in the code. By doing so, the
stopping time is overestimated by a factor of (1 + ), where is the
dust-to-gas ratio. The value of remains always below unity in all
our simulations. Therefore, despite using this approximation, we
obtain meaningful results for the dust evolution in the disc (at least
for the short evolutionary times considered). The dust behaviour as
a function of the Stokes number is discussed in Section 4.1.
We chose the grain size for which the particles have a Stokes
number close to unity. For the parameters considered in Sect. 2.2.1,
this size corresponds to s = 1 mm. The reason for that choice, is
that we wish to study the particles that: i) feel the strongest radial
drift, and ii) concentrate the most efficiently in the pressure maxima
of the CBD.
2.3 Set of simulations
The simulations are divided into two sets according to their ec-
centricity: eB = 0.50 and eB = 0.75. For each set, the incli-
nations are divided in prograde cases with iB = {0◦, 30◦, 60◦},
the polar case with iB = 90◦, and the retrograde cases with
iB = {120◦, 150◦, 180◦}. The names of the simulations are listed
in Table 1. In Figure 1, we show the binary eccentricity and orien-
tation with respect to the circumbinary disc. In the following, the
disc is always seen face-on with the binary inclined inside the inner
cavity.
3 RESULTS
Figures 2 and 3 show the gas and dust surface density maps --
along with the dust-to-gas ratio -- for eB = 0.50 and eB =
0.75 (respectively). The CBD is shown after 50 binary orbits. At
this evolutionary stage, the structures in the dust distribution have
reached a quasi steady-state.
We observe features of different kind in the CBD. A horse-
shoe, defined as the main gas over-density at the edge of the disc
cavity. A dust ring along the disc inner edge, caused by the radial
drift of the dust particles. When the horseshoe traps most of the
dust, a large clump of dust forms on top of it. We also observe that
the dust can also get trapped in other regions along the dust ring.
Remarkably, these small clumps are not bound to the horseshoe. In
the following, we describe each of these disc features: as a function
of iB (Sect. 3.1) and eB (Sect. 3.2).
2 we adopt an intrinsic dust grain density equal to 3 gr cm−3, a typical
value used for astrophysical silicates.
MNRAS 000, 1 -- ?? (2019)
Dusty clumps in circumbinary discs
3
3.1 From prograde to retrograde cases
Prograde cases: the most striking structure observed in Figs. 2
and 3 is the horseshoe at the inner edge of the disc. This is seen
only at inclinations iB = {0°, 30°}, and for both eccentricities.
This has already been reported in previous works of coplanar black
hole binaries (Shi et al. 2012; D'Orazio et al. 2013; Farris et al.
2014) and stellar binaries (Ragusa et al. 2017). In particular, in the
latter the authors report a dust concentration at the location of the
horseshoe. This structure is comparable to the large dust clump ob-
served in e50-i0, e50-i30, e75-i0, and e75-i30. However, we also
observe the formation of small dust clumps outside the horseshoe.
The differences between large and small clumps are their angular
size and density. Quantitatively, the large clump is more than five
times denser compared to the average dust density along the dusty
ring; while the small clumps are only twice denser compared to the
average value (see Figure 4). In addition, the large clump tends to
overlap with the horseshoe covering roughly 60° in azimuth; while
small clumps cover smaller azimuthal sectors (less than 30°). Re-
markably, for iB = 60°, both the horseshoe and the large clump
disappear. Instead, several small clumps appear along the dust ring.
When this happens, the binary-triggered spirals and streamers are
the only gas structures observed.
The cases with iB = {0°, 30°} show a annular-shaped feature just
outside the dense inner dust ring. This is easily seen in the dust-
to-gas ratio maps of Figs. 2 and 3. This structure forms due to the
action of the gaseous spiral arms on the dust, which modify the gas
density in that region -- and hence the Stokes number. This speeds
up the radial velocity of the dust particles, which eventually leads
to the formation of a dusty gap in the disc.
Polar configuration: for iB = 90°, we see a set of 5 small clumps
evenly distributed along the dust ring (Fig. 4). The density of each
of these clumps is about twice the average density of the dust ring.
Retrograde cases: in this configuration, we observe a remarkable
difference between eB = 0.5 and eB = 0.75, as explained in Sec-
tion 3.2. Nevertheless, a common aspect is that the densest gas re-
gions are displaced toward the companion's orbital apoastron. This
is shown in the bottom rows of Figs. 2 and 3. The location of the
densest region can be explained by the binary perturbations on the
gas disc: the secondary star acts braking the surrounding gas. The
deceleration is higher in the disc region closer to the companion
(Nixon et al. 2011).
Additionally, we note that the cavity size decreases with in-
creasing binary inclination as found by Miranda & Lai (2015). For
instance, this effect can be easily seen by considering the shape of
the dust ring. In addition, the cavity becomes eccentric and its cen-
tre does not match with the centre of mass of the system (Dunhill
et al. 2015). The coplanar cases (iB = 0°) present cavity sizes in
agreement with the classic result of (Artymowicz & Lubow 1994)
for different eccentricities.
We observe comparable values of the dust-to-gas ratio ( =
ρd/ρg) in the large and small clumps. This is because for the large
clump, the dust and gas densities are high; while for the small
clump both densities are lower. The implications of the high dust-
to-gas ratio values will be discussed in Section 4.1.3.
Finally, we observe the formation of a circumprimary disc for
iB = 120° and iB = 150°. These kind of discs are likely transient
and are hardly seen at this numerical resolution. This is because the
low density of particles inside the cavity translates into a high nu-
merical viscosity. Therefore, the circumstellar discs quickly drain
into the stars, which are modelled as sink particles as in Price et al.
(2018b).
4
P.P. Poblete, N. Cuello & J. Cuadra
Figure 2. Gas (first column), dust (middle column) surface density in gr/cm2 and the dust-to-gas ratio in last column, after 50 binary orbits at eB = 0.50.
From upper to bottom are the different inclinations, iB = {0°, 30°, 60°, 90°, 120°, 150°, 180°} respectively. The gray circle on the bottom-right corner of
the two first columns represents the Gaussian kernel (5 au in diameter) used to smooth the images. This size is consistent with the highest ALMA angular
resolution reached so far. We recall that in our simulations the physical quantities are computed using the smoothing length.
MNRAS 000, 1 -- ?? (2019)
Dusty clumps in circumbinary discs
5
Figure 3. Same as Fig. 2, but for the case eB = 0.75.
MNRAS 000, 1 -- ?? (2019)
6
P.P. Poblete, N. Cuello & J. Cuadra
Figure 4. Normalised dust surface density along the dust ring. Prograde
(iB = 0°), polar (iB = 90°) and retrograde (iB = 180°) cases are shown
in red, blue, and green (respectively). Solid and dashed lines correspond to
eB = 0.50 and eB = 0.75, respectively.
3.2 Eccentricity 0.50 versus 0.75
We find that the disc cavities are larger for eB = 0.75 compared
to eB = 0.5, for the very same inclination. This is in agreement
with Miranda & Lai (2015). Also, the higher the binary eccentric-
ity, the higher the density of the gas spirals and streamers. This is
well seen for regrades cases with eB = 0.75: the gas disc exhibits
both prominent spiral structures and multiple spiral arms. The latter
are concentrated in a specific azimuthal sector of the CBD. These
disc features are not observed for eB ≤ 0.50, neither for coplanar
configurations as in Ragusa et al. (2017). In Section 4.1, we show
why the small dust clumps only form in retrograde configurations
for eB = 0.75.
4 DISCUSSION
4.1 Clumps
As reported in Sect. 3, large and small clumps can form along
the dust ring according to the binary parameters. The mechanism
of dust trapping by a local azimuthal gas over-density -- namely
the horseshoe -- has already been studied extensively in previous
works (e.g., Johansen et al. 2004; Birnstiel et al. 2013; Owen &
Kollmeier 2017; Ragusa et al. 2017). However, to the best of our
knowledge, the formation of small dust clumps in CBDs has not
been reported yet. These features are particularly prominent for
iB = 90°, but they also appear for iB = {0°, 30°}.
The main characteristic of the small clumps is that, although
they form on top of local gas over-densities, they do not necessarily
follow the gas. This is in contrast to the large clump, explained
above. Below, we focus in more detail on the formation of small
clumps.
4.1.1 Formation of a single small clump
For the coplanar retrograde (iB = 180°) case, small clumps are
only observed for eB = 0.75 (as opposed to eB = 0.50). This
is because a higher eccentricity favours the formation of more
prominent gas spirals and denser streamers. These gaseous features
caused by the inner binary are crucial to trigger the small clump
formation. Specifically, the binary-induced gaseous streams perturb
the dusty ring through aerodynamical drag. The strength of the lat-
ter heavily depends on the Stokes number (see Eq. (1)). For sake
Figure 5. Sketch of the mechanism of formation of a small clump. The
gaseous spirals are represented in blue and the dust ring in red. The forming
small clump is highlighted in black. The disc rotation and the binary centre
of mass location are shown with green arrows. The length of the orange
arrows indicates the magnitude of radial drift, which is higher in the region
between the spirals.
of simplicity, let's assume that the dust ring has a constant density,
which is a reasonable approximation before any clump forms along
the ring. Since the grain size is fixed, then the Stokes number only
depends on the gas density.
The process of small clump formation is shown in Figure 5
where we schematically represent the motion of the dust on top of
the gaseous spirals. The inner spiral (called the head) is caused by
the secondary star, while the outer one (called the tail) corresponds
to the spiral formed in the previous orbit. The tail is at a larger
distance from the binary compared to the head. The dust ring (in
red) is deformed by the two gaseous spirals. The Stokes number in
both spirals is less than one due to their high gas density, whereas
it is higher in the region between the two spirals. For our disc pa-
rameters, 1 mm grains have St << 1 in the spirals and St ∼ 1
in between. Due to the strong coupling, the inner dust ring follows
the head, whereas the outer dust ring follows the tail. In addition,
the bending of the dust ring generates a significant radial density
gradient. As a consequence, the dust particles in between the spi-
rals radially drift towards the head. This effect is the strongest for
mm-sized grains because their Stokes number is close to one (Wei-
denschilling 1977). Therefore, millimetric dust is efficiently accu-
mulated at the head location, where a small clump begins to form.
4.1.2 Evolution and behaviour of a single small clump
Because of the periodic perturbation of the disc caused by the bi-
nary, gas spirals continuously form at a specific azimuthal sector
of the disc. Therefore, after one binary orbit, the previous head be-
comes the tail (with a small clump attached to it) and the innermost
gas stream becomes the head. This explains why dust clumps are
periodically formed in the CBD in our models.
Once the clumps form, there are two possible dynamical out-
comes: they can either be disrupted or keep growing. The survival
and long-term behaviour of these individual structures are key for
grain growth, and consequently for planetesimal formation in the
disc. The survival of the small clumps is related to the local Stokes
number.
The region where gaseous spirals are formed has high density
-- clumps will have Stokes numbers less than one in there. There-
fore, even though the small clump formation happens in that re-
MNRAS 000, 1 -- ?? (2019)
gion, the clump can also be easily stretched and potentially dis-
rupted. Such stretching is caused by the gradient of angular veloc-
ity within the clump due to interaction between the clump and the
spiral. More specifically, the head of the spiral moves faster com-
pared to the regions behind it. To characterise the evolution of the
small clumps, we define their corresponding survival timescale as
the time from their formation until their disruption. In all our simu-
lations, we observe survival timescales of at least one orbital period
at the clump radial distance. It is precisely when the clump com-
pletes the first orbit and returns to the spiral-forming region that it
can be potentially disrupted. We also note that the inclination af-
fects the clump survival time. For instance, in cases with iB = 90°,
the binary torque does not strongly affect the azimuthal velocity
gradient of the gas spiral. In this configuration, the small clumps
survive for several tens of orbits.
Besides disruption, the clump can also be fed after complet-
ing an orbit. Figure 6 shows all the possible scenarios that a small
clump can experience. The first one, called A, happens when a dust
stream falls exactly onto the small clump, making it grow. The other
three scenarios (B, C, and D) lead to clump disruption. Once the
small clump is disrupted, its remnants are later fed by the outer
dust stream. However, it is worth noting that the previous clump
never reforms as such. To sum up, it is possible to either generate a
more massive clump (A); or to disrupt the main clump generating
several ones (B, C, and D).
Throughout all the simulation, the circumbinary disc is period-
ically perturbed by the secondary star. This ensures the continuous
formation of clumps as previously described. At the end of our sim-
ulations (i.e. after 100 binary orbits) the dust disc exhibits a similar
morphology as the one observed after 50 binary orbits. Neverthe-
less, it is worth noting that the individual small clumps shown at 50
orbits are not necessarily the same as the ones present at the end of
the simulation.
4.1.3 Dust-to-gas ratio
Interestingly, if the dust-to-gas ratio becomes close to one then self-
induced dust traps (Gonzalez et al. 2017) could appear in the disc.
The increase of the dust-to-gas ratio could also potentially trig-
ger the streaming instability (Johansen & Youdin 2007). Therefore
these could be sweet spots for grain growth and planetesimal for-
mation in the CBD. However, these dynamical effects are not seen
in our simulations due to their short evolutionary time. In addition,
SPH is not the most suitable method to capture streaming instabil-
ity effects due to the two-fluid numerical scheme (Laibe & Price
2012a). We did not run the models for longer, as the density ap-
proximation made in Sect. 2.2.2, namely ρ = ρg, becomes less
valid precisely as the dust-to-gas ratio increases.
4.2 Clumps formation with different grain sizes
So far, we have only discussed the large and small clump formation
for one specific grain size, namely mm-sized particles. Here, we
present simulations with two other grain sizes: 100 µm and 1 cm.
Figures 7 and 8 show the dust morphology for e50-i0 and e50-i90,
respectively. In addition, we also show a test simulation without
aerodynamic drag (labelled "no drag"), where the dust particles be-
have as test particles.
Comparing Figures 7 and 8, we see that the grain size plays
a crucial role in the formation of clumps. This is because the dust
coupling (i.e. the Stokes number) depends linearly on the grain size,
MNRAS 000, 1 -- ?? (2019)
Dusty clumps in circumbinary discs
7
Figure 6. The four possible scenarios that can experience a small clump
after its formation. These examples are taken from the simulation e50-i90
at different times. In A, the clump is fed by a dust stream, without being
disrupted. In B, we see a third clump forming in between a disrupted clump.
In C and D, the dust streams feeds the back and front side (respectively) of
a disrupted clump.
as shown in Eq. 1. In particular, the proposed formation mechanism
(see Sect. 4.1.1) is the most efficient for dust grains with a Stokes
number varying from one to slightly less than one along the orbit,
which corresponds to 1 mm for the disc parameters we have chosen.
Nevertheless, for other disc parameters (e.g. disc mass, temperature
and density profiles, etc.), the condition St ∼ 1 would correspond
to a different grain size, which would form structures similar to the
ones reported here.
Interestingly, the dust ring and the clumps are mainly caused
by gas drag effects, and not only by the binary gravitational pertur-
bations. For instance, in the no-drag simulations we do not observe
any clumps or dusty rings along the cavity.
4.3 Dust features as indicators of unseen stellar companion
Direct observations of stellar companions in binary systems are
particularly challenging. Specifically, there are strong limitations
to properly resolve the separation between two stars. This is even
worse if the companion is less massive and therefore fainter com-
pared to the main star. However, here we have shown that some
prominent disc features can be triggered by the gravitational in-
teraction of a low-mass stellar companion inside the cavity. More
specifically, the set of simulations of this work explore a modest but
meaningful region of the vast space of parameters, namely the bi-
nary eccentricity and inclination. Hence, in principle, the disc fea-
tures could be used to infer the orbit of a potentially unseen stellar
companion.
4.3.1 Remarkable dust structures in CBDs
Large dusty clump within a gas horseshoe. These features appear
in all our simulations with iB ≤ 30°. Both have the same properties
as the ones reported by Ragusa et al. (2017). They are not produced
by a vortex, and have a high contrast compared to the rest of the
disc. Thus, a large dusty clump on top of a gas horseshoe could be
an indicator of an inner companion with an orientation close to the
disc plane.
Embedded small clumps in a dusty ring. For all our highly-
inclined simulations (60° ≤ iB ≤ 120°), we observed several
small dusty clumps embedded in the disc. Interestingly, in the polar
case, clumps are azimuthally equidistant between them. Therefore,
several small clumps along the dust ring could indicate the presence
of a highly inclined inner companion. Note that the case e75-i180
8
P.P. Poblete, N. Cuello & J. Cuadra
shows a clump-ring structure too, therefore a highly eccentric and
retrograde companion is also able to create the same feature. In
Appendix A we show that an inner planet-mass companion does
not produce such structures, which allows us to set a lower mass
threshold for structure formation in the CBD.
Smooth dust ring. All the cases that do not show any remarkable
features (horseshoe or clumps) in the dust ring are included in this
category. In the absence of structure it is hard to draw any conclu-
sion on whether there is a single star or a binary system. Never-
theless, the inner cavity structure could help to infer the presence
of an inner companion. For instance, during the early disc evolu-
tion, a large inner cavity of several tens of au strongly suggests the
presence of a binary system inside the cavity. However, for more
evolved discs and if no accretion is detected, the cavity is more
likely to be caused by photoevaporative processes (Alexander et al.
2006; Owen 2016)
Caution is required when interpreting our results since this
analysis mainly applies to grains with a Stokes number close to
unity. See for instance the structures obtained for different grain
sizes in Figures 7 and 8. HD 142527 is a notorious example where
two different dust structures coexist: a large clump at millimetric
wavelengths (Boehler et al. 2018) and small clumps at centimetric
wavelengths (Casassus et al. 2015b).
When spirals are observed, their morphology and their az-
imuthal concentration in particular can provide further information
on their dynamical origin. Besides an inner binary, flybys (Cuello
et al. 2019a), planets (Dong et al. 2015), self-gravitating discs (Dip-
ierro et al. 2015a; Forgan et al. 2018), or shadows (Montesinos et al.
2016; Montesinos & Cuello 2018; Cuello et al. 2019b) can also pro-
duce spiral arms in the disc. The main difference is that the spirals
caused by an eccentric inner companion are often multiple and well
concentrated in one azimuthal sector of the disc, as opposed to the
other mechanisms.
4.3.2 The AB Aurigae system
AB Aurigae -- an Herbig Ae star of the A0 spectral type and mass
2.4± 0.2 M(cid:12) (DeWarf et al. 2003) -- exhibits a very complex mor-
phology, both in gas and dust: i) multiple spiral arms in scattered
light (Fukagawa et al. 2004; Corder et al. 2005; Hashimoto et al.
2011), ii) a horseshoe-shaped dust trap (Tang et al. 2012; Pacheco-
Vázquez et al. 2016), iii) and a large dust cavity at a distance from
the star between ∼ 70 and 100 au (Hashimoto et al. 2011; Tang
et al. 2012). A single planet has been proposed to explain some of
the observed features (Hashimoto et al. 2011; Fuente et al. 2017;
Tang et al. 2017). Tang et al. (2012) were only able to explain the
cavity size by adding a body at r ∼ 45 au and M = 0.03M(cid:12);
whereas Fuente et al. (2017) managed to explain the emission of
the dust disc by putting a Jupiter-mass planet at r = 94 au. It is
however challenging to explain all the aforementioned features si-
multaneously.
Instead, the binary scenario proposed by Price et al. (2018b)
for HD 142527 seems more promising. As a matter of fact, there is
a striking similarity between the structures observed in AB Aurigae
and those in HD 142527. Pirzkal et al. (1997) gives an upper limit
of 0.25 M(cid:12) down to 60 au for an possible inner stellar companion
in AB Aurigae. It is worth noting that mass constraint at small radii
is difficult to quantify, the mass upper limit could be greater. There-
fore, a low mass ratio binary scenario for AB Aurigae is reasonable.
Based on the disc observations, our models suggest the presence of
an inner stellar companion -- undetected so far.
Figure 7. Dust distribution of the case e50-i0 after 50 initial binary orbits,
for three sizes of dust grains, plus one simulation without the aerodynamic
drag. The dust grain sizes are 100 microns, 1 millimetre, and 1 centimetre.
The no-drag simulation was made for s = 1 mm.
As mentioned in Sect 4.3.1, the absence of a gas horseshoe
and the multiple spiral arms suggest a high eccentricity and an in-
clination higher than 30°. In particular, the case e50-i60 reproduces
the observed dust distribution remarkably well (see Figure 9), si-
multaneously explaining the multiple and azimuthally concentrated
gaseous spirals in the disc (not shown). Figure 9 shows a com-
parison between the observed dust continuum emission at 1.3 mm
(Tang et al. 2012) and the dust distribution in e50-i60, where the
dust over-density is seen as a large clump due to beaming effects.
In addition, Rivière-Marichalar et al. (2019) very recently reported
the observation of clumps in HCN -- a good tracer of cold and dense
gas -- around the inner edge of the disc. This supports the idea that
the dusty clumps might be real. It is worth to mention however that
a bad coverage of the uv plane could produce artificial clumps in the
reconstructed intensity map. Indeed, the 0.9 mm image presented
by Tang et al. (2017) shows a continuous inner ring, rather than
clumps. Future observations at a higher angular resolution and with
better uv plane coverage are required in order to reveal whether
small clumps are indeed embedded in the disc of AB Aur.
In summary, our results strongly motivate the search for a stel-
lar companion within the cavity of the disc around AB Aurigae.
More specifically, an i) unequal-mass, ii) eccentric, and iii) inclined
stellar binary can potentially explain most (if not all) the observed
disc features.
5 CONCLUSIONS
We performed 3D SPH gas and dust simulations of circumbi-
nary discs (CBDs) around binaries with different eccentricities
and inclinations. We considered unequal-mass binaries similar to
HD 142527. This allowed us to characterise the disc morphology
for different combinations of orbital parameters of the companion.
The main conclusions of our work are the following:
MNRAS 000, 1 -- ?? (2019)
Dusty clumps in circumbinary discs
9
planets (also known as polar Tatooines). In this regard, systems
similar to e50-i90 are better candidates to host this type of polar
planets.
Interestingly, the asymmetric disc features aforementioned
can be much more easily detected than the binary itself. Consider-
ing the systems categorised as Giant Discs by Garufi et al. (2018),
several of them show multiple, asymmetric, relatively faint arm-
like structures on large scales. Based on our results, we are inclined
to think that there might be binaries in the cavities of several of
those discs. In particular, we strongly suggest the presence of an
eccentric and inclined inner companion in AB Aurigae.
ACKNOWLEDGEMENTS
We thank the anonymous referee for valuable comments and sug-
gestions that have improved our work. Figures 2, 3, 6, 7, 8, 9(b), A1
and B1 were made with SPLASH (Price 2007). The Geryon2 clus-
ter housed at the Centro de Astro-Ingenieria UC was used for the
calculations performed in this paper. The BASAL PFB-06 CATA,
Anillo ACT-86, FONDEQUIP AIC- 57, and QUIMAL 130008 pro-
vided funding for several improvements to the Geryon/Geryon2
cluster. The authors acknowledge support from CONICYT project
Basal AFB-170002. PPP and JC acknowledge support from Inicia-
tiva Científica Milenio via the Núcleo Milenio de Formación Plan-
etaria. NC acknowledges financial support provided by FONDE-
CYT grant 3170680. This project has received funding from the Eu-
ropean Union's Horizon 2020 research and innovation programme
under the Marie Skłodowska-Curie grant agreement No 823823.
REFERENCES
ALMA Partnership et al., 2015, ApJ, 808, L3
Alexander R. D., Clarke C. J., Pringle J. E., 2006, MNRAS, 369, 216
Aly H., Dehnen W., Nixon C., King A., 2015, MNRAS, 449, 65
Andrews S. M., et al., 2018, ApJ, 869, L41
Armitage P. J., 2018, A Brief Overview of Planet Formation. p. 135,
doi:10.1007/978-3-319-55333-7_135
Artymowicz P., Lubow S. H., 1994, ApJ, 421, 651
Ataiee S., Pinilla P., Zsom A., Dullemond C. P., Dominik C., Ghanbari J.,
2013, A&A, 553, L3
Avenhaus H., Quanz S. P., Schmid H. M., Meyer M. R., Garufi A., Wolf S.,
Dominik C., 2014, ApJ, 781, 87
Avenhaus H., et al., 2018, ApJ, 863, 44
Bate M. R., Bonnell I. A., Price N. M., 1995, MNRAS, 277, 362
Biller B., et al., 2012, ApJ, 753, L38
Birnstiel T., Dullemond C. P., Pinilla P., 2013, A&A, 550, L8
Boehler Y., Weaver E., Isella A., Ricci L., Grady C., Carpenter J., Perez L.,
2017, ApJ, 840, 60
Boehler Y., et al., 2018, ApJ, 853, 162
Bromley B. C., Kenyon S. J., 2015, ApJ, 806, 98
Casassus S., et al., 2013, Nature, 493, 191
Casassus S., et al., 2015a, ApJ, 811, 92
Casassus S., et al., 2015b, ApJ, 812, 126
Christiaens V., Casassus S., Perez S., van der Plas G., Ménard F., 2014,
ApJL, 785, L12
Christiaens V., et al., 2018, A&A, 617, A37
Claudi R., et al., 2019, A&A, 622, A96
Corder S., Eisner J., Sargent A., 2005, ApJ, 622, L133
Cuello N., Giuppone C. A., 2019, arXiv e-prints, p. arXiv:1906.10579
Cuello N., et al., 2019a, MNRAS, 483, 4114
Cuello N., Montesinos M., Stammler S. M., Louvet F., Cuadra J., 2019b,
A&A, 622, A43
D'Orazio D. J., Haiman Z., MacFadyen A., 2013, MNRAS, 436, 2997
Figure 8. Same as Fig. 7, but for the case e50-i90.
(i) An inner stellar companion with a low inclination (iB ≤ 30°)
with respect to the CBD is able to trigger a horseshoe-like structure
in both the gas and the dust. Additionally, small dust clumps can
also appear along the dusty ring. The latter had not been reported
by previous works.
(ii) For an inner stellar companion on a highly inclined orbit
with respect to the CBD (60° ≤ iB ≤ 120°) the dust ring breaks
into small clumps, evenly distributed along the cavity.
(iii) For high eccentricities (eB = 0.75), the binary perturba-
tions on the gas disc become stronger -- especially for retrograde
cases (120° ≤ iB ≤ 180°). This translates into denser gas struc-
tures: spirals, streams, and horseshoes. Therefore, the higher the
eccentricity the easier the formation of clumps.
(iv) The small clump structure strongly depends on the Stokes
number of the dusty ring. The formation mechanism is most effi-
cient when the Stokes number is close to unity. A detailed descrip-
tion of its formation and evolution is given in Section 4.1.
Circumbinary discs are often thought to be unfavourable sys-
tems for planetesimal formation, because of the high relative ve-
locities expected among solid bodies (Bromley & Kenyon 2015). In
this work, we have found that high dust-to-gas ratio clumps form in
the inner regions of discs around unequal-mass, eccentric, and in-
clined binaries (especially for polar configurations). Such clumps
could then constitute sweet spots for dust accumulation and grain
growth (Gonzalez et al. 2017; Owen & Kollmeier 2017), suggest-
ing that CBDs could potentially be efficient planetesimal cradles.
Within this context, polar circumbinary discs are of particular
interest since we have shown they form stable and prominent dusty
clumps (e50-i90 and e75-i90). Theoretical models first predicted
the existence of this kind of polar discs (Aly et al. 2015; Martin
& Lubow 2017; Zanazzi & Lai 2018; Lubow & Martin 2018), as
the one very recently discovered in HD 98800 by Kennedy et al.
(2019). It seems reasonable to think that planets will eventually
form in these polar circumbinary discs. Based on that assumption,
Cuello & Giuppone (2019) showed that binaries with mild eccen-
tricities are more likely to retain their circumbinary P-type polar
MNRAS 000, 1 -- ?? (2019)
10
P.P. Poblete, N. Cuello & J. Cuadra
Figure 9. Comparison between the observation at 1.3 mm of AB Aurigae (left column) and the dust distribution in e50-i60 after 100 orbits (right column).
(a): Dust continuum emission at 1.3 mm. The black cross represents the stellar peak. The blue and red crosses mark the peak of 12CO J=2→1: highest
blue-shifted and red-shifted peak respectively. (b): Surface density map from e50-i60 convolved by a 50 au beam (shown in the left corner), consistent with
the observations in (a). (c): 1.3 mm intensity along the dust ring. (d): Surface density along the dust ring normalised to the average dusty ring density. The
frame orientation is chosen so the dust over-density is roughly at the same azimuthal position as that in the the observation, and inclined −23° with respect to
the x-axis as the observed system Tang et al. (2012). The θPA in (c) and (d) represents the offset of the PA position at 121.3°, measured from the north in a
clockwise sense. The observations in (a) and (c) are taken from Figures 1 and 11 in Tang, A&A, 547, A84, 2012, reproduced with permission © ESO.
DeWarf L. E., Sepinsky J. F., Guinan E. F., Ribas I., Nadalin I., 2003, ApJ,
590, 357
Dipierro G., Pinilla P., Lodato G., Testi L., 2015a, MNRAS, 451, 974
Dipierro G., Price D., Laibe G., Hirsh K., Cerioli A., Lodato G., 2015b,
Monthly Notices of the Royal Astronomical Society: Letters, 453, L73
Dipierro G., Laibe G., Price D. J., Lodato G., 2016, Monthly Notices of the
Royal Astronomical Society: Letters, 459, L1
Dong R., Zhu Z., Rafikov R. R., Stone J. M., 2015, ApJ, 809, L5
Dunhill A. C., Cuadra J., Dougados C., 2015, MNRAS, 448, 3545
Epstein P. S., 1924, Physical Review, 23, 710
Farris B. D., Duffell P., MacFadyen A. I., Haiman Z., 2014, ApJ, 783, 134
Forgan D. H., Ilee J. D., Meru F., 2018, ApJ, 860, L5
Fuente A., et al., 2017, ApJ, 846, L3
Fukagawa M., et al., 2004, ApJ, 605, L53
Fukagawa M., Tamura M., Itoh Y., Kudo T., Imaeda Y., Oasa Y., Hayashi
S. S., Hayashi M., 2006, ApJ, 636, L153
Gaia Collaboration et al., 2016, A&A, 595, A1
Garufi A., et al., 2018, A&A, 620, A94
Gonzalez J. F., Laibe G., Maddison S. T., 2017, MNRAS, 467, 1984
Hashimoto J., et al., 2011, ApJ, 729, L17
Johansen A., Youdin A., 2007, ApJ, 662, 627
Johansen A., Andersen A. C., Brandenburg A., 2004, A&A, 417, 361
Kennedy G. M., et al., 2019, Nature Astronomy, p. 189
Lacour S., et al., 2016, A&A, 590, A90
Laibe G., Price D. J., 2012a, MNRAS, 420, 2345
Laibe G., Price D. J., 2012b, MNRAS, 420, 2365
Lubow S. H., Martin R. G., 2018, MNRAS, 473, 3733
Lyra W., Lin M.-K., 2013, ApJ, 775, 17
Marino S., Perez S., Casassus S., 2015, ApJL, 798, L44
Martin R. G., Lubow S. H., 2017, ApJ, 835, L28
Meheut H., Meliani Z., Varniere P., Benz W., 2012, A&A, 545, A134
Miranda R., Lai D., 2015, MNRAS, 452, 2396
MNRAS 000, 1 -- ?? (2019)
Montesinos M., Cuello N., 2018, MNRAS, 475, L35
Montesinos M., Perez S., Casassus S., Marino S., Cuadra J., Christiaens V.,
2016, ApJ, 823, L8
Nakagawa Y., Sekiya M., Hayashi C., 1986, Icarus, 67, 375
Nixon C. J., Cossins P. J., King A. R., Pringle J. E., 2011, MNRAS, 412,
1591
Nixon C., King A., Price D., 2013, MNRAS, 434, 1946
Owen J. E., 2016, Publications of the Astronomical Society of Australia,
33, e005
Owen J. E., Kollmeier J. A., 2017, MNRAS, 467, 3379
Pacheco-Vázquez S., et al., 2016, A&A, 589, A60
Perez S., et al., 2015, ApJ, 798, 85
Pinilla P., et al., 2018, ApJ, 859, 32
Pirzkal N., Spillar E. J., Dyck H. M., 1997, ApJ, 481, 392
Price D. J., 2007, Publications of the Astronomical Society of Australia, 24,
159
Price D. J., et al., 2018a, Publications of the Astronomical Society of Aus-
tralia, 35, e031
Price D. J., et al., 2018b, MNRAS, 477, 1270
Raghavan D., et al., 2010, The Astrophysical Journal Supplement Series,
190, 1
Ragusa E., Dipierro G., Lodato G., Laibe G., Price D. J., 2017, Monthly
Notices of the Royal Astronomical Society, 464, 1449
Rivière-Marichalar P., Fuente A., Baruteau C., Neri R., Treviño-Morales
S. P., Carmona A., Agúndez M., Bachiller R., 2019, ApJ, 879, L14
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Shi J.-M., Krolik J. H., Lubow S. H., Hawley J. F., 2012, ApJ, 749, 118
Tang Y. W., Guilloteau S., Piétu V., Dutrey A., Ohashi N., Ho P. T. P., 2012,
A&A, 547, A84
Tang Y.-W., et al., 2017, ApJ, 840, 32
Tokovinin A., 2014, AJ, 147, 86
Weidenschilling S. J., 1977, MNRAS, 180, 57
Zanazzi J. J., Lai D., 2018, MNRAS, 473, 603
van der Marel N., Cazzoletti P., Pinilla P., Garufi A., 2016, ApJ, 832, 178
APPENDIX A: PLANETARY COMPANION IN THE
POLAR CASE
To test whether the structures we found in the dust disc can also be
triggered by planetary-mass companions, we performed two simu-
lations with the very same setup as e50-i90, but with a companion
mass of 10 MJ. In addition, we considered different values for the
initial disc inner edge: Rin = 90 au and Rin = 60 au. Given the re-
duced strength of the gravitational perturbations, the reduced cavity
size allows us to bring material closer to the inner planetary com-
panion. In Figure A1, we show the dust and gas distributions after
50 planetary orbits. Regardless of the value of Rin, the structures
are different from the ones obtained for a stellar companion (see
middle row in Fig. 2). The only remarkable feature is the smooth
dust ring. No small clumps nor spirals are observed. In summary,
the latter features can only be triggered by a stellar companion for
the configuration considered.
APPENDIX B: NUMERICAL TESTS FOR SMALL CLUMP
FORMATION
In order to test whether or not the small dusty clumps reported in
this work were caused by numerical effects, we performed a con-
vergence test for different resolutions in dust. This was done for the
case e50-i90 (i.e. our more representative example) at three resolu-
tions: 1.25· 104, 105 (this work), and 4· 105 dust particles; keeping
the gas resolution fixed to 106 particles. These three simulations
MNRAS 000, 1 -- ?? (2019)
Dusty clumps in circumbinary discs
11
Figure A1. Gas (top panels) and dust (bottom panels) morphology for the
case e50-i90 after 50 binary orbits, but with a companion mass reduced to
10 MJ (i.e. 50 times less massive). The initial disc inner edge is set at 90 au
and 60 au in the left and right panels, respectively. A planetary companion
is not able to trigger the formation of dusty clumps in the disc.
are shown in the left, middle, and right panels of Figure B1 (respec-
tively). We observe that the low and high resolution tests exhibit the
same structures (dust ring plus small clumps) as the simulation with
105 dust particles. Their azimuthal positions are identical for the
resolutions considered. We also see that the proposed mechanism
to form small clumps by bending the dust ring (see Sect. 4.1.1) still
holds -- regardless of the number of dust particles. The five clumps
seen in Fig. 2 do not appear here because of the earlier evolutionary
stage of the disc (11 binary orbits instead of 50). Since we observe
emergent small clumps, it is reasonable to expect these features to
appear eventually. Based on these results, we thus conclude that
the formation of small dusty clumps is a physical process, which is
properly captured at the resolution of 105 dust particles.
Finally, in order to test whether the formation of small dusty
clumps is affected by our approximation of the Stokes number (see
Eq. 1) we performed a shorter simulation without the approxima-
tion, i.e., computing St ∝ ρ
−1
T = (ρg + ρd)−1. We did this for
e50-i60 because it is the case that shows the highest dust-to-gas
ratio values. Hence, it presents the most significant difference be-
tween the Stokes number computed with and without the approxi-
mation. Figure B2 shows the azimuthal density profile of both mod-
els. We see that regardless of the way we compute the Stokes num-
ber, the small clumps form, and also do it at approximately the same
azimuthal position. We therefore conclude that the approximation
does not affect our results in any significant way.
APPENDIX C: DISC VERTICAL SCALE-HEIGHT AT
DIFFERENT RADIAL DISTANCES
In Figure C1 we show < h > /H for our simulations with eB = 0.5
(left panel) and with eB = 0.75 (right panel). Since < h > /H < 1,
the disc is properly resolved in the vertical direction. The value of
the Shakura -- Sunyaev viscosity αSS can be easily inferred from the
values of < h > /H in Figure C1. In this case, it is of the order
12
P.P. Poblete, N. Cuello & J. Cuadra
Figure B1. Dust distribution for the case e50-i90 after 11 binary orbital periods for different resolutions in dust: 1.25 · 104 (left), 105 (middle), and 4 · 105
(right) SPH dust particles. The gas resolution is fixed to 106 SPH gas particles for all the simulations. The formation of dusty clumps is not affected by the
dust resolution.
Figure C1. Radial profiles of < h > /H after 50 binary orbits for all the
simulations with eB = 0.50 (left panel) and with eB = 0.75 (right panel).
Figure B2. Azimuthal density profile along the dust ring after 21 binary
orbits. The blue line corresponds to e50-i60 with the Stokes number com-
puted with the approximation (St ∝ ρ−1
g ). Instead, the orange line corre-
sponds to e50-i60 without the approximation (St ∝ (ρg + ρd)−1, where
ρg and ρd are the gas and dust density respectively).
of 0.005 as mentioned in the text and as in the simulations in Price
et al. (2018b).
MNRAS 000, 1 -- ?? (2019)
|
1505.06204 | 1 | 1505 | 2015-05-22T20:00:29 | Formation of planetary debris discs around white dwarfs II: Shrinking extremely eccentric collisionless rings | [
"astro-ph.EP",
"astro-ph.SR"
] | The formation channel of the tens of compact debris discs which orbit white dwarfs (WDs) at a distance of one Solar radius remains unknown. Asteroids that survive the giant branch stellar phases beyond a few au are assumed to be dynamically thrust towards the WD and tidally disrupted within its Roche radius, generating extremely eccentric (e>0.98) rings. Here, we establish that WD radiation compresses and circularizes the orbits of super-micron to cm-sized ring constituents to entirely within the WD's Roche radius. We derive a closed algebraic formula which well-approximates the shrinking time as a function of WD cooling age, the physical properties of the star and the physical and orbital properties of the ring particles. The shrinking timescale increases with both particle size and cooling age, yielding age-dependent WD debris disc size distributions. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 8 (2015)
Printed 28 August 2018
(MN LATEX style file v2.2)
Formation of planetary debris discs around white dwarfs II:
Shrinking extremely eccentric collisionless rings
Dimitri Veras1⋆, Zoe M. Leinhardt2, Siegfried Eggl3, Boris T. Gansicke1
1Department of Physics, University of Warwick, Coventry CV4 7AL, UK
2The School of Physics, University of Bristol, Bristol BS8 1TL, UK
3IMCCE Observatroire de Paris, UPMC, 77 Av. Denfert-Rochereau, 75014 Paris, France
Accepted 2015 May 22. Received 2015 May 20; in original form 2015 February 25
ABSTRACT
The formation channel of the tens of compact debris discs which orbit white dwarfs (WDs)
at a distance of one Solar radius remains unknown. Asteroids that survive the giant branch
stellar phases beyond a few au are assumed to be dynamically thrust towards the WD and
tidally disrupted within its Roche radius, generating extremely eccentric (e > 0.98) rings.
Here, we establish that WD radiation compresses and circularizes the orbits of super-micron
to cm-sized ring constituents to entirely within the WD's Roche radius. We derive a closed
algebraic formula which well-approximates the shrinking time as a function of WD cooling
age, the physical properties of the star and the physical and orbital properties of the ring
particles. The shrinking timescale increases with both particle size and cooling age, yielding
age-dependent WD debris disc size distributions.
Key words: minor planets, asteroids: general -- stars: white dwarfs -- methods: numerical --
celestial mechanics -- planet and satellites: dynamical evolution and stability -- protoplanetary
discs
1 INTRODUCTION
Mounting discoveries of debris orbiting white dwarfs (WDs)
presage increased scrutiny of post-main-sequence planetary
systems. The over 30 dusty discs
(Zuckerman & Becklin
1987; Becklin et al. 2005; Kilic et al. 2005; Reach et al. 2005;
Farihi et al. 2009; Barber et al. 2014; Bergfors et al. 2014;
Rocchetto et al. 2015) and 7 gaseous discs (Gansicke et al. 2006,
2007, 2008; Gansicke 2011; Farihi et al. 2012; Melis et al. 2012;
Wilson et al. 2014; Manser et al. 2015) so far detected all have
radial extents of just about 1 Solar radius ≈ 0.005 au < 106
km. Such compact configurations are absent from main sequence
planetary studies because they cannot exist; the discs would be
inside of the star! Consequently, achieving an understanding of
how these post-main-sequence discs are formed requires a dif-
ferent approach.
Protoplanetary discs orbiting young main-sequence stars
form out of a collapsing stellar birth cloud, but WD discs
must instead be formed from the tidal disruption of objects
which encounter the WD (Graham et al. 1990; Jura 2003;
Debes et al. 2012; Bear & Soker 2013). However, because of
mass loss and stellar tides on the giant branch phases of stel-
lar evolution, there exists a "planet desert". Mass loss serves
⋆ E-mail:[email protected]
c(cid:13) 2015 RAS
to push individual planets outward (sometimes to the interstel-
lar medium; Veras et al. 2011; Veras & Tout 2012; Adams et al.
2013; Veras et al. 2014a), and may cause multiple planets to be-
come unstable (Duncan & Lissauer 1998; Debes & Sigurdsson
2002; Veras et al. 2013a; Voyatzis et al. 2013; Mustill et al.
2014; Veras & Gansicke 2015), leading to collisions, engulfment
within the star, or escape from the system. Concurrently, tidal
effects between the giant star and a close-in planet could over-
come the outward evolution due to mass loss and cause the star
to swallow the planet (Villaver & Livio 2009; Kunitomo et al.
2011; Mustill & Villaver 2012; Spiegel & Madhusudhan 2012;
Adams & Bloch 2013; Villaver et al. 2014). Therefore, we ex-
pect a void of planetary bodies within a few au of WDs.
Asteroids, comets, moons and planets are instead likely to
exist beyond a few au in WD systems. Remnant exo-asteroid or
exo-Kuiper belts can then dynamically interact with these plan-
ets, flinging asteroids towards the WD and eventual destruc-
tion (Bonsor et al. 2011; Debes et al. 2012; Frewen & Hansen
2014)1. The disruption of a single asteroid, which must be on
a highly eccentric orbit in order to enter the disruption sphere
1 Other potential reservoirs of disrupted material, such as exo-Oort
clouds, are both compositionally (Zuckerman et al. 2007) and dynam-
ically (Veras et al. 2014b; Stone et al. 2015) unlikely to represent the
primary source of the discs.
2
Veras, Leinhardt, Eggl & Gansicke
(or Roche radius) of the WD, results in a highly eccentric ring
(Veras et al. 2014c, hereafter Paper I). By modelling the aster-
oid as a rubble pile composed of equal-mass, equal-radii inde-
structible spheres, Paper I found that the resulting eccentric ring
is collisionless. The tidal disruption of multiple asteroids in quick
succession might instead produce collisional rings or discs. How-
ever, here we treat the single asteroid (collisionless) case as the
next logical step to follow Paper I. The resulting eccentric ring
maintains the same orbit as the progenitor, but precesses due to
general relativity. Paper I found that the eccentricity e of these
orbits always satisfies e > 0.98, with typical values of e ∼ 0.995;
an asteroid with a semimajor axis a of 10 au that skims a typical
WD would satisfy e ≈ 0.99999. If the rings remain collisionless,
then without the influence of additional forces, they will never
change shape.
However, observations of gaseous WD discs reveal circular
or near-circular geometries. Double-peaked metal line emission
profiles suggest e ≈ 0.02 and e ∼ 0.2 in the two most robustly-
constrained systems (Gansicke et al. 2006, 2008). The process
of converting a highly eccentric orbit into a nearly circular one,
while drastically reducing the semimajor axis, has rarely been
discussed in the context of WDs, and is the issue we address in
this paper.
To do so, we introduce radiative forces into the two-body
problem, as each ring constituent can be considered to be orbit-
ing the WD unperturbed by any other body if the constituents
are indeed collisionless. We focus on radiation and do not con-
sider effects from tidal dissipation in this work. Regardless, radi-
ation forces may dominate over tidal dissipation effects because
the latter are likely to cause orbital changes of a ring particle
on a much longer timescale. The effects of tidal dissipation in
small and light ring constituents should be negligible, as can
be deduced directly from the equations of motion (e.g. equation
8 of Beaug´e & Nesvorn´y 2012) or from relations for the orbit-
averaged semimajor axis and eccentricity damping timescales
(e.g. equations 10-11 of Nagasawa & Ida 2011). Both of those
studies account for dynamical tides, a physical description in
which highly eccentric orbits may be treated.
Radiation,
in the restricted, specific case of Poynting-
Robertson drag, has become a crucial component of explain-
ing accretion from compact circumstellar discs onto WDs
(Bochkarev & Rafikov 2011; Rafikov 2011a,b; Metzger et al.
2012). Wyatt et al. (2014) has constrained the possible size dis-
tributions of the accreting material by fitting model param-
eters to observationally-inferred accretion rate distributions.
We know that accretion onto the WD occurs because associ-
ated with every dusty and gaseous disc are unmistakable sig-
natures of photospheric metal pollution. These metals cannot
arise from stellar dredge-up nor from the interstellar medium
(ISM) due to the high accretion rate, the distribution of inter-
stellar clouds and the existence of metal-rich WDs that lack hy-
drogen (Aannestad et al. 1993; Friedrich et al. 2004; Jura 2006;
Kilic & Redfield 2007; Gansicke et al. 2008; Farihi et al. 2010).
Here, we consider WD radiation in a different, earlier context,
before a circular compact disc has formed. We are particu-
larly motivated by the suggestion that Poynting-Robertson drag
plays an important role in the dynamics of highly-eccentric or-
bits (Vokrouhlick´y et al. 2001), as the tidally-disrupted debris
from Paper I settled into such orbits.
In Section 2, we present our framework of perturbations
due to WD radiation. Sections 3 and 4 describe how radiation
changes orbits on short and long timescales, respectively. Based
on these results, we derive the key formulae for orbital shrinkage
in Section 5 before discussing the implications for the resulting
ring structure due to collisions and the Yarkovsky effect in Sec-
tion 6 and concluding in Section 7.
2 GENERAL FORMULATION
We let ~r(t) represent the distance from the centre of the ring
particle to the centre of the WD, ~v(t) the velocity of the par-
ticle with respect to the centre of the WD, m the particle's
mass, A the particle's momentum-carrying cross-sectional area,
c the speed of light and L(t) the time-varying luminosity of the
WD. Then the extra acceleration of the particle due to stellar
radiation is (Veras et al. 2015a)
=
(1)
AL
4πmcr2"QabsI + QrefI + kQyarY#~ι,
(cid:18) d~v
dt(cid:19)rad
where Qabs is the absorption efficiency, Qref is the reflection ef-
ficiency, Qyar = Qabs − Qref , I is the 3x3 unit matrix, and Y is
the 3x3 Yarkovsky matrix (given in equation 28 of Veras et al.
2015a). The thermal redistribution parameter k satisfies 0 6
k 6 1/4. The three terms in equation (1) refer to the contri-
butions from radiation which is absorbed, directly reflected and
thermally emitted. The vector
~ι ≡(cid:18)1 −
~v · ~r
cr (cid:19) ~r
r −
~v
c
(2)
represents the relativistically-corrected direction of incoming ra-
diation.
The above equations hold when the particle is at least a
few orders of magnitude larger than the wavelength of incoming
radiation, ¯λ. For WDs, ¯λ changes with cooling age (time since
the star became a WD). Young (∼4 Myr old), very hot WDs
with temperatures of about 40000 K will have a value of ¯λ that
is just under 1 × 10−7 m. Alternatively, old (∼2 Gyr old) and
cold (6000 K) WDs have ¯λ ≈ 5×10−7 m. Effectively, throughout
a WD's lifetime, ¯λ varies by less than an order of magnitude,
and is always safely smaller than a micron.
The contribution from the Yarkovsky term is negligible
when the particle size is larger than about 1 cm - 10 m
(Veras et al. 2015a) and when the particle spins, such that some
of the absorbed radiation is redistributed before being emit-
ted. When the Yarkovsky effect is active, it can represent the
dominant contribution to the motion: the resulting perturbation
may eject an asteroid orbiting a 103L⊙ evolved star (such as an
asymptotic giant branch star or very young WD) in just tens of
million years. Here, we assume that the tidal disruption of an
asteroid into sub-metre-sized pieces occurs on a much shorter
timescale.
Paper I suggests the number of orbits required for the as-
teroids that were modeled to dissociate completely into its in-
dividual constituents was tens to hundreds. The pericentres of
these asteroid orbits were within about 10% of the WD's Roche
radius, and the modeled asteroids were strengthless rubble piles.
If the asteroids were instead internally more cohesive, then the
disruption limits would change. Asteroids closer to the outer
edge of the Roche radius may indeed disrupt into sufficiently
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
small bits on longer timescales, such that the Yarkovsky effect
does play a significant role.
We discuss this possibility further in Section 6. However, for
the body of this paper, we continue the story begun in Paper I
and consider the subsequent evolution of the rings formed from
disruption due to close pericentre passages. We assume that the
constituents of these rings are small enough to be unaffected by
the Yarkovsky effect, and we operate on this assumption for the
remainder of Sections 2-5. Consequently, equation (1) simplifies
to
(cid:18) d~v
dt(cid:19)Y=0
rad
where
=
ALQ
4πmcr2~ι.
Q ≡ Qabs + Qref .
(3)
(4)
The luminosity of the WD is a steep function of the star's
cooling age within the first hundreds of Myr. We adopt the same
dependence on cooling age as in equation (6) of Bonsor & Wyatt
(2010) and re-express their equation as
L(t) = 3.26L⊙(cid:18) M⋆
0.6M⊙(cid:19)(cid:18)0.1 +
t
Myr(cid:19)−1.18
,
(5)
which is applicable for about the first 9 Gyr of WD cooling
(sufficient for our purposes). Here M⋆ refers to the stellar mass.
We now determine how the radiative force from equation (3)
reshapes the debris rings. First, we consider the perturbations
on orbital timescales, and then model the cumulative (i.e. long-
term, averaged, or secular) consequences.
3 UNAVERAGED EQUATIONS OF MOTION
The orbit of the particle and star is fixed unless acted upon by
perturbations, such as that from equation (3). Based on equa-
tions from Efroimsky (2005) and Gurfil (2007), Veras & Evans
(2013a) outlined an algorithm which can produce expressions in
orbital elements for the time evolution of a (semimajor axis), e
(eccentricity), i (inclination), Ω (longitude of ascending node),
ω, (argument of pericentre), f (true anomaly), and the pericen-
tre q = a(1 − e) and apocentre q = a(1 + e), in the perturbed
two-body problem. This technique eliminates all Cartesian com-
ponents, and does not make any assumptions about the orbital
elements except that the motion remains bounded. This facet
is important because the eccentricities we will be treating are
near unity, and hence traditional disturbing functions expanded
about small eccentricities would not be applicable2.
The complete equations of motion due to the perturba-
tion from equation (3) have already been derived in Veras et al.
(2015a), and we do not repeat them here. One quantity of par-
ticular interest not emphasized in that paper is the closeness of
the orbital pericentre to the WD for an extremely eccentric or-
bit. We obtain the time evolution of q here by manipulating the
equations for the evolution of a and e in Veras et al. (2015a),
and find
2 Veras (2007) showed how increasing the order of the expansion helps
only until the Sundman criterion (Ferraz-Mello 1994) is violated.
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
Forming WD discs II: shrinking
3
= −
(cid:18) dq
dt(cid:19)Y=0
sin2(cid:0) f
+
2πm (1 + e)2
×"
ALQ (1 + e cos f )2
c2(cid:19)#.
2(cid:1) [2 − e − e cos f ]
(cid:18) 1
a (1 − e2)
sin f
2na2√1 − e2(cid:18) 1
c(cid:19)
(6)
We need to know how much radiation will drag a parti-
cle located at the pericentre of the orbit (f = 0) towards (or
away from) the WD relative to the particle's unperturbed oscu-
lating location. In the next section we can obtain an order-of-
magnitude estimate of the maximum value of this displacement
by averaging equation (6) over an entire orbit.
4 AVERAGED EQUATIONS OF MOTION
In order to consider the cumulative effect of WD radiation over
many orbits, we analyze the averaged equations of motion. We
perform the averaging over the true anomaly such that for an
arbitrary variable β,
We obtain (Veras et al. 2015a)
0
1
(cid:28) dβ
dt(cid:29) ≡
2πZ 2π
dt(cid:19)Y=0+ = −
*(cid:18) da
*(cid:18) de
dt(cid:19)Y=0+ = −
dβ
(1 + e cos f )2 df.
dt (cid:0)1 − e2(cid:1)3/2
ALQ(cid:0)2 + 3e2(cid:1)
4πmac2 (1 − e2)3/2 ,
8πma2c2√1 − e2
5ALQe
.
(7)
(8)
(9)
Equations (8) and (9) agree with Wyatt & Whipple (1950), and
demonstrate that radiation decreases both the semimajor axis
and eccentricity over time. Note that the (1/c) terms vanish
upon averaging.
There is no long-term change in the argument or longitude
of pericentre
*(cid:18) dω
dt(cid:19)Y=0+ =*(cid:18) d
dt (cid:19)Y=0+ = 0
(10)
which importantly illustrates that the long-term precession of
the orbit is dictated by general relativity (in the absence of
other, additional forces). Finally, we find that
dt(cid:19)Y=0+ = −
*(cid:18) dq
*(cid:18) deq
dt(cid:19)Y=0+ = −
ALQ√1 − e2 (4 − e)
8πmac2 (1 + e)2
ALQ√1 − e2 (4 + e)
8πmac2 (1 − e)2
,
,
(11)
(12)
proving that Poynting-Robertson drag decreases both the peri-
centre and apocentre over time.
A closer look at equation (11) can reveal by how much radi-
ation causes the pericentre to shrink per orbit. Let ∆ represent
the actual close encounter distance minus the distance predicted
by Newtonian gravity with no additional forces per orbit. We
find
∆ ≈ −0.364 m(cid:18)Q
√1 − e2 (4 − e)
(1 + e)2
10 au(cid:17)1/2(cid:18) m
(cid:19)(cid:16) a
1012 kg(cid:19)−
1
3
4
Veras, Leinhardt, Eggl & Gansicke
×(cid:18) M⋆
0.6M⊙(cid:19)−
1
2(cid:18)
ρ
2 g/cm3(cid:19)−
2
3(cid:18)0.1 +
t
Myr(cid:19)−1.18
(13)
where ρ represents the particle density. Hence, along a single
orbit, radiation from a newly-born WD drags a body towards
the star by about 4 m at the pericentre. For cooling ages of 1
Myr and 1 Gyr, this value drops to about 0.4 m and 0.4 mm. For
highly eccentric orbits, these estimates would be reduced by at
least one order of magnitude. To place these values in context,
consider that a typical WD radius is ≃ 0.015R⊙ = 104 km
(Hamada & Salpeter 1961; Holberg et al. 2012; Parsons et al.
2012) and the maximum possible extent of the WD disruption
sphere is ≃ 3.6R⊙ = 2.5 × 106 km (Paper I). Hence, over the
lifetime of the Universe, this gradual pericentre drift is negligible
compared to the WD radius or its disruption sphere.
Further, we know that general relativity produces no such
accumulation of inward or outward drag despite causing a po-
tentially relatively large deviation from the Newtonian orbit at
each pericentre passage. The maximum value of this deviation
is about 9 km ×(M⋆/M⊙), and the body is pushed towards the
star only when e . 0.359 (Veras 2014).
By using both the averaged equations of motion (equations
8, 9 and 10) and a prescription for the luminosity evolution
(equation 5), we can numerically determine the long-term or-
bital evolution of particles from WD radiation. This approach
requires solving coupled differential equations, as we cannot find
an explicit solution of the general equations. However, in the
specific case of high eccentricity, we have found an explicit so-
lution, as described in the next section.
First, we numerically integrate these full averaged equations
of motion (equations 5, 8, 9 and 10) to determine the evolution
for WD planetary systems. Results of this integration are plot-
ted as solid lines in Figure 1 for bodies with radii (R) spanning
four orders of magnitude. Only for this size range (10−5 − 10−1
m) can we be sure that other radiative effects do not come into
play. Each panel illustrates a different WD cooling age tini; older
WDs take longer to shrink orbits. For simplicity, we adopted
a characteristic WD mass (M⋆) of 0.6M⊙, characteristic body
density (ρ) of 2 g/cm3 and Q = 1, for the integrations. As-
suming Q = 1 means that no directed scattering or reemission
occurs (i.e. Qref = Qyar = 0), such that all radiation is absorbed
and emitted omnidirectionally. This assumption accurately re-
produces the properties of small carbonaceous dust particles
with albedos of about 0.005. Alternatively, typical asteroids with
albedos between 0.1 and 0.3 would satisfy 1.1 < Q < 1.3.3
We also adopted one of the pericentre values used in Paper
I to better demonstrate the link with that paper and to pro-
vide analytically-motivated initial parameters. In that respect,
qmin = 0.0126R⊙ = 5.86 × 10−5 au is the radius of a typical
0.6M⊙ WD, and qmax = 2.73R⊙ = 0.013 au is the maximum
value of the disruption sphere for that WD mass.
The figure demonstrates that for a wide range of WD cool-
ing ages, all particles in this size range will maintain their orig-
inal semimajor axes for the vast majority of their evolution,
before their orbits suddenly shrink to within the WD Roche ra-
dius. Also plotted is the apocentre of the orbit. For the majority
of the particle's evolution, the particle maintains its initial ec-
centricity; circularization occurs only as the semimajor axis is
shrinking drastically. This behaviour is mirrored for particle evo-
lution at later WD cooling ages, although the collision timescale
is increased, as will be shown in the next section.
5 HIGH ECCENTRICITY LIMIT
Numerical simulations from the last subsection demonstrate
that a particle on an initially highly eccentric orbit will remain
so for the vast majority of its lifetime. This result suggests that
considering the high-eccentricity limit of the equations might
represent a good approximation to the actual motion. Our ob-
jective in this subsection is to express this motion by an explicit
algebraic expression, which will facilitate future study.
We perform series expansions about ǫ = 1− e, where ǫ ≪ 1,
and denote the result with a tilde. We obtain
5ALQ
8√2πmac2 (1 − e)3/2
^*(cid:18) da
dt(cid:19)Y=0+ = −
32√2πmac2 (1 − e)1/2 + O(cid:16)(1 − e)1/2(cid:17) ,
dt(cid:19)Y=0+ = −
^*(cid:18) de
9ALQ
+
8√2πma2c2 (1 − e)1/2 + O(cid:16)(1 − e)1/2(cid:17) .
5ALQ
(14)
Now suppose the debris ring is formed at the WD cooling age
tini, such that at this time each particle adopts semimajor axis
and eccentricity values of aini and eini. We see from equation
(11) that
(15)
e→1*(cid:18) dq
dt(cid:19)Y=0+ = 0.
lim
(16)
The semimajor axis and eccentricity change due to radiative
effects, but on average the pericentre remains constant. The
eccentricity and semimajor axis evolution can thus be approxi-
mated as
a(t) [1 − e(t)] ≈ aini (1 − eini) .
Equation (17) may also be obtained by simultaneously solving
the leading terms present in equations (14) and (15). By using
the relation in equation (17), we obtain
(17)
^*(cid:18) da
dt(cid:19)Y=0+ ≈ −
8√2πma3/2
5ALQ√a
ini c2 (1 − eini)3/2
.
(18)
In order to solve equation (18) algebraically and explicitly
for the semimajor axis, we express L as in equation (5) but
replace the power law exponent of −1.18 with −6/5, such that
(19)
L ≈ L0(cid:18) 1
10
+
t
Myr(cid:19)−6/5
3 The albedo distribution for main belt asteroids is bimodal, with
peaks at about 0.05 and 0.25.
where L0 = 3.26L⊙(M/0.6M⊙) and is hence constant. Also, we
express A and m in terms of the body's radius (R) and density
(ρ), as we are more interested in the dependence on size than
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
Forming WD discs II: shrinking
5
a(t) = aini − K + K2
4aini
.
(21)
Further, because the pericentre is constant, the apocentre reads
q(t) ≈ a(t) − 2qini.
(22)
We plot the curves predicted from equation (21) as dashed
lines on Fig. 1. The agreement with the true solution is excellent,
for both the semimajor axis and apocentre, and for all cooling
ages relevant to WD pollution, at least to the precision we seek.
Only physical solutions from equation (21) are plotted. Af-
ter a body encounters the Roche radius, the time evolution is
stopped. Otherwise, the semimajor axis would then unphysically
increase after achieving a minimum. To stop the time evolution
at the appropriate time, one can multiply equation (21) by the
appropriate Heaviside function.
Of particular interest is the shrinking time, tshr, which we
define as the time taken for the particle's initial orbit to be com-
pressed entirely within the WD Roche radius. By using equation
(21), we find
10 (tshr + tini)
Myr
10tini
=(cid:18)1 +
(cid:18)1 +
−(cid:18) 64 · 23/10π
(cid:19)−1/5
L0QMyr!(cid:16)√aini ±p2qini + RRoche(cid:17) (23)
75 · 51/5 (cid:19) c2ρRq3/2
Myr (cid:19)−1/5
ini
where RRoche is the Roche radius of the WD. From this equa-
tion tshr may be solved for explicitly; we adopted the (physical)
solution with the lower sign.
Figure 1. Orbit shrinkage of various particle sizes (with radii R)
from WD radiation. The three panels refer to different WD ages
(tini) for a WD with a typical mass of 0.6M⊙. From top to bottom,
these ages correspond to luminosities of about 22L⊙, 2 × 10−3L⊙ and
1 × 10−4L⊙. Plotted in each pair of similarly-coloured curves are the
evolution of the semimajor axis (a; lower curve) and apocentre (q;
upper curve). The eccentricity of the orbit becomes zero only when
those two lines converge. Solid lines are the result of numerical in-
tegrations of the full equation of motion (8-9) and the dashed lines
are from our analytical approximations (equations 21-22). The ini-
tial semimajor axis was taken to be aini = 10 au and the pericentre
to be qini = qmin + 5%(qmax − qmin), which corresponds to the or-
bital parameters of some rings formed in Paper I. All orbits become
compressed to entirely within the Roche radius.
on mass. The final result can be expressed compactly through
the following auxiliary variable
K ≡(cid:18) 75 · 51/5
×"(cid:18)1 +
32 · 23/10π(cid:19)
Myr (cid:19)−1/5
10tini
L0QMyr
c2ρRaini (1 − eini)3/2!
Myr(cid:19)−1/5# .
−(cid:18)1 +
10t
The evolution of the semimajor axis is
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
(20)
We use equation (23) to generate Figure 2, which illustrates
the dependencies of tshr on R and a. For observational perspec-
tive we display the range of cooling ages of known WD with de-
bris discs (≈ 30 Myr - 1.5 Gyr; Farihi et al. 2009; Girven et al.
2012; Bergfors et al. 2014) with a thick brown horizontal bar and
vertical brown lines. The steep increase in shrinking timescale
during the first Gyr of WD evolution is a reflection of the star's
rapidly dwindling luminosity at early cooling ages. Neverthe-
less, for all cooling ages, the radiation is still strong enough to
shrink and circularize the orbits on timescales orders of mag-
nitude smaller than the WD cooling age. Coincidentally, the
largest particles for which the Yarkovsky effect play no role
(∼ 10−1 m) also approach the maximum particle size at which
tshr ≪ tini, at least for aini = 10 au. Indeed, we do not yet have
any observational constraints on aini; equations such as (23) may
help motivate likely values.
6 DISCUSSION
With our results, we can now try to construct a consistent pic-
ture from Paper I about how tidal disruption of asteroids is
followed by orbit compression. A single asteroid is tidally dis-
rupted into a collisionless ring with an unknown size distribu-
tion. The orbit of the ring is equivalent to the orbit of the pro-
genitor, whose eccentricity exceeds 0.98 and whose semimajor
axis exceeds a few au. A Main Belt-like asteroid would satisfy
aini ≈ 5 au, a Kuiper Belt-like asteroid would satisfy aini ≈ 30
au, and an asteroid from a debris belt in a system like Fomal-
haut would satisfy aini ≈ 100 au. Because the ring is assumed to
be collisionless, the bodies will not gravitationally perturb one
6
Veras, Leinhardt, Eggl & Gansicke
Figure 2. The shrinking timescale tshr as a function of WD cooling
age tini for curves of different particle radii (upper panel) and initial
semimajor axes (lower panel). In all cases, a ρ = 2 g/cm3 particle
orbits a 0.6M⊙ WD. In the upper panel, we adopt aini = 10 au
and qini = qmin + 10%(qmax − qmin) (solid lines) and qini = qmin +
5%(qmax − qmin) (dashed lines). The similarly-coloured pairs of lines
each corresponds to a different order of magnitude of particle radius.
In the lower panel, all particles have R = 10−3 m and qini = qmin +
5%(qmax − qmin). Marked for reference as a thick brown bar and
vertical brown lines is the range of cooling ages at which WD discs
have been observed. All curves in both panels demonstrate that the
shrinking timescales are orders of magnitude shorter than the cooling
times.
another. Every body will experience radiation forces in the form
of Poynting-Robertson drag.
6.1 Collisions upon contraction
If the ring contains differently-sized bodies, then the effects of
Poynting-Robertson-drag alone will cause the orbits of these
bodies to contract at different rates. As smaller bodies contract
more quickly, their pericentres will precess more rapidly due to
general relativity. Consequently, collisions might occur. Consider
equation 19 of Paper I:
Pω ≈ 0.15 Myr(cid:20) 1 − e2
1 − 0.9992(cid:21)(cid:18) MWD
0.6M⊙(cid:19)−3/2(cid:16) a
1 au(cid:17)5/2
(24)
which illustrates that the general relativistic precession period
Pω ∝ a5/2 and is also a function of e and MWD. Both these last
two quantities remain nearly constant as the orbit contracts by
a factor of tens to hundreds4. Therefore, suppose an asteroid on
a a = 10 au orbit is broken up into large and small particles with
radii Rlarge and Rsmall. The larger particle ring will continue to
precess with a period of about 47 Myr while the smaller particle
ring is shrunk; at a = 1 au, that small particle ring's precession
period is just about 0.15 Myr. Consequently, when tshr & Pω, we
can expect collisions to be significant; for small enough particles
such that tshr ≪ Pω, collisional effects would be negligible.
6.2 Yarkovsky perturbations on chunky rings
If there exist ring constituents that remain greater than about
one metre in diameter over long-enough timescales, then the
Yarkovsky effect will likely become the dominant contribution
to the motion. We mentioned this possibility in Section 2, but
now provide some quantification.
Let Q ≡ QabsI + QrefI + kQyarY. The components of Q are
a function of spin state and orbital position, velocity and time.
Although the simplified spherical internal heat model provided
by Broz (2006) allows the time evolution of the semimajor axis
or eccentricity to be expressed (see Sections 2.3.2 and 2.3.3 of
Broz 2006, or Vokrouhlick´y 1998, Vokrouhlick´y & Farinella 1998
or Vokrouhlick´y 1999), that model itself might not be applicable
for the extremely high eccentricities considered in this paper. For
a particle on an eccentric orbit, the seasonal component of the
Yarkovsky effect contains many forcing periods; the amplitudes
of those frequencies are functions of the orbital eccentricity.
Regardless of these difficulties, we now provide a zeroth-
order demonstration of the potential importance of the effect.
Each element of Q is strictly bounded between 0 and 2, allow-
ing us to bound the extent of the potential perturbation. By
assuming that the matrix elements are independent of position
and velocity, Veras et al. (2015a) obtained the resulting orbital
element-based unaveraged equations, as well as the averaged
equations to leading order in (1/c). By expanding their equa-
tions A3 and A4 about ǫ = 1 − e, ǫ ≪ 1, we obtain
Comparing equations (25)-(26) with (14)-(15) reveals that (14)-
(15) are a factor of about 1/c smaller than (25)-(26), and can
4 Eventually, the eccentricity starts to decrease appreciably, poten-
tially increasing the factor in square brackets by up to a factor of
about 500. Nevertheless, this eccentricity decrease does not occur un-
til the particle has migrated nearly all the way to the Roche radius.
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
2 (1 − e)
1
AL
^(cid:28) da
dt(cid:29) =
×
^(cid:28) de
dt(cid:29) =
×
4πcmna2(cid:20)
·
Q12 − Q21
Q23 − Q32
Q13 − Q31
·
8πcmna3
Q12 − Q21
Q23 − Q32
Q13 − Q31
AL
+
1
4(cid:21)
cos i
sin i sin Ω
sin i cos Ω
cos i
sin i sin Ω
sin i cos Ω
+ O (1 − e) ,
+ O(cid:0)√1 − e(cid:1) .
(25)
(26)
be neglected when the Yarkovsky acceleration is "turned on".
Further, because the leading-order eccentricity term of equation
(25) is of order 1/(1 − e) whereas that of equation (26) is of
order of just unity, the Yarkovsky acceleration might cause the
orbit to shrink quickly at a nearly fixed eccentricity.
The off-diagonal terms will determine the sign of the expres-
sions, and hence whether the orbit shrinks or expands, circular-
izes or becomes parabolic. In fact, because the initial eccentricity
is so high, on average we might expect about half of the large
chunks in the ring to be ejected. Given that the Yarkovsky ac-
celeration affects particles only above a certain size, the extra
movement of these particles with respect to the relative velocity
of the constituents after break-up might lead to collisions, which
could produce a damping effect.
Although we wish to obtain an expression similar to equa-
tion (21) for the Yarkovsky effect, we cannot do so, but can
make some progress towards a solution. In principle, we could
follow the same procedure as in Section 5 to obtain an approx-
imate closed formula for a(t) due to the leading-order terms in
equations (25)-(26). In the high eccentricity limit, we can relate
a and e to one another and their values at a particular cooling
age because of the similar forms of equations (25)-(26). Dividing
equation (25) by equation (26) and integrating yields
2a2 = 2a2
ini +
1
2
(e − eini) + ln(cid:18) 1 − eini
1 − e (cid:19).
(27)
However, we cannot go further and substitute equation (27) into
equation (25) and then integrate primarily because equations
(25)-(26) depend on the inclination and longitude of ascending
node (which are time-dependent) and the elements of Q (which
may be time-dependent).
7 CONCLUSIONS
We have addressed a problem in the formation of discs which
orbit WDs at a distance of 1R⊙: how to shrink the 10 − 100
au-scale extremely eccentric (e > 0.98) bound orbits of debris
from tidally-disrupted asteroids. We demonstrated that this or-
bit compression readily occurs for particle sizes of 10−5 − 10−1
m on timescales many orders of magnitude shorter than the WD
cooling age, due to WD radiation alone. We provided explicit ap-
proximations for the shrinking time (equation 23) as well as for
the semimajor axis and apocentre evolution (equations 21-22)
during the orbit compression (the pericentre remains constant).
After the debris is perturbed close to or within the Roche ra-
dius, subsequent evolution is likely to be dictated by collisional
and sublimative forces.
ACKNOWLEDGMENTS
We thank the referee, Miroslav Broz, for reading our paper care-
fully and providing detailed comments, which have helped us
improve and tighten the manuscript. We also thank J.J. Her-
mes for useful discussions. The research leading to these re-
sults has received funding from the European Research Council
under the European Union's Seventh Framework Programme
(FP/2007-2013) / ERC Grant Agreements n. 320964 (WD-
Tracer) and n. 282703 (NEOShield), as well as from Paris Obser-
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
Forming WD discs II: shrinking
7
vatory's ESTERS (Environment Spatial de la Terre: Rechreche
& Surveilance) travel grants.
REFERENCES
Aannestad, P. A., Kenyon, S. J., Hammond, G. L., & Sion,
E. M. 1993, AJ, 105, 1033
Adams, F. C., Anderson, K. R., & Bloch, A. M. 2013, MNRAS,
432, 438
Adams, F. C., & Bloch, A. M. 2013, ApJL, 777, L30
Barber, S. D., Kilic, M., Brown, W. R., & Gianninas, A. 2014,
ApJ, 786, 77
Beaug´e, C., & Nesvorn´y, D. 2012, ApJ, 751, 119
Bear, E., & Soker, N. 2013, New Astronomy, 19, 56
Becklin, E. E., Farihi, J., Jura, M., et al. 2005, ApJL, 632, L119
Bergfors, C., Farihi, J., Dufour, P., & Rocchetto, M. 2014, MN-
RAS, 444, 2147
Bochkarev, K. V., & Rafikov, R. R. 2011, ApJ, 741, 36
Bonsor, A., & Wyatt, M. 2010, MNRAS, 409, 1631
Bonsor, A., Mustill, A. J., & Wyatt, M. C. 2011, MNRAS, 414,
930
Broz, M. 2006, Ph.D. Thesis
Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1
Debes, J. H., & Sigurdsson, S. 2002, ApJ, 572, 556
Debes, J. H., Walsh, K. J., & Stark, C. 2012, ApJ, 747, 148
Duncan, M. J., & Lissauer, J. J. 1998, Icarus, 134, 303
Efroimsky, M. 2005, Annals of the New York Academy of Sci-
ences, 1065, 346
Farihi, J., Jura, M., & Zuckerman, B. 2009, ApJ, 694, 805
Farihi, J., Barstow, M. A., Redfield, S., Dufour, P., & Hambly,
N. C. 2010, MNRAS, 404, 2123
Farihi, J., Gansicke, B. T., Steele, P. R., et al. 2012, MNRAS,
421, 1635
Farihi, J., Gansicke, B. T., & Koester, D. 2013, Science, 342,
218
Ferraz-Mello, S. 1994, Celestial Mechanics and Dynamical As-
tronomy, 58, 37
Frewen, S. F. N., & Hansen, B. M. S. 2014, MNRAS, 439, 2442
Friedrich, S., Jordan, S., & Koester, D. 2004, A&A, 424, 665
Gansicke, B. T., Marsh, T. R., Southworth, J., & Rebassa-
Mansergas, A. 2006, Science, 314, 1908
Gansicke, B. T., Marsh, T. R., & Southworth, J. 2007, MN-
RAS, 380, L35
Gansicke, B. T., Koester, D., Marsh, T. R., Rebassa-
Mansergas, A., & Southworth, J. 2008, MNRAS, 391, L103
Gansicke, B. T. 2011, American Institute of Physics Conference
Series, 1331, 211
Girven, J., Brinkworth, C. S., Farihi, J., et al. 2012, ApJ, 749,
154
Graham, J. R., Matthews, K., Neugebauer, G., & Soifer, B. T.
1990, ApJ, 357, 216
Gurfil, P. 2007, Acta Astronautica, 60, 61
Hamada, T., & Salpeter, E. E. 1961, ApJ, 134, 683
Holberg, J. B., Oswalt, T. D., & Barstow, M. A. 2012, AJ, 143,
68
Jura, M. 2003, ApJL, 584, L91
Jura, M. 2006, ApJ, 653, 613
Kilic, M., von Hippel, T., Leggett, S. K., & Winget, D. E. 2005,
ApJL, 632, L115
8
Veras, Leinhardt, Eggl & Gansicke
Kilic, M., & Redfield, S. 2007, ApJ, 660, 641
Kunitomo, M., Ikoma, M., Sato, B., Katsuta, Y., & Ida, S.
2011, ApJ, 737, 66
Manser, C.J., et al. MNRAS In Prep
Melis, C., Dufour, P., Farihi, J., et al. 2012, ApJL, 751, L4
Metzger, B. D., Rafikov, R. R., & Bochkarev, K. V. 2012, MN-
RAS, 423, 505
Mustill, A. J., & Villaver, E. 2012, ApJ, 761, 121
Mustill, A. J., Veras, D., & Villaver, E. 2014, MNRAS, 437,
1404
Nagasawa, M., & Ida, S. 2011, ApJ, 742, 72
Parsons, S. G., Marsh, T. R., Gansicke, B. T., et al. 2012,
MNRAS, 420, 3281
Rafikov, R. R. 2011a, ApJL, 732, L3
Rafikov, R. R. 2011b, MNRAS, 416, L55
Rafikov, R. R., & Garmilla, J. A. 2012, ApJ, 760, 123
Reach, W. T., Kuchner, M. J., von Hippel, T., et al. 2005,
ApJL, 635, L161
Rocchetto, M., Farihi, J., Gansicke, B. T., & Bergfors, C. 2015,
MNRAS, 449, 574
Spiegel, D. S., & Madhusudhan, N. 2012, ApJ, 756, 132
Stone, N., Metzger, B. D., & Loeb, A. 2015, MNRAS, 448, 188
Veras, D. 2007, Celestial Mechanics and Dynamical Astronomy,
99, 197
Veras, D., Wyatt, M. C., Mustill, A. J., Bonsor, A., & Eldridge,
J. J. 2011, MNRAS, 417, 2104
Veras, D., & Tout, C. A. 2012, MNRAS, 422, 1648
Veras, D., Mustill, A. J., Bonsor, A., & Wyatt, M. C. 2013a,
MNRAS, 431, 1686
Veras, D., Hadjidemetriou, J. D., & Tout, C. A. 2013b, MN-
RAS, 435, 2416
Veras, D., & Evans, N. W. 2013a, Celestial Mechanics and
Dynamical Astronomy, 115, 123
Veras, D., & Evans, N. W. 2013b, MNRAS, 430, 403
Veras, D. 2014, MNRAS, 442, L71
Veras, D., Evans, N. W., Wyatt, M. C., & Tout, C. A. 2014a,
MNRAS, 437, 1127
Veras, D., Shannon, A., Gansicke, B. T. 2014b, MNRAS, 445,
4175
Veras, D., Leinhardt, N. W., Bonsor, A., Gansicke, B. T. 2014c,
MNRAS, 445, 2244
Veras, D., Jacobson, S. A., Gansicke, B. T. 2014d, MNRAS,
445, 2794
Veras, D., Gansicke, B. T. 2015, MNRAS, 447, 1049
Veras, D., Eggl, S., Gansicke, B. T., 2015, MNRAS, In Press,
arXiv:1505.01851
Villaver, E., & Livio, M. 2009, ApJL, 705, L81
Villaver, E., Livio, M., Mustill, A. J., & Siess, L. 2014, ApJ,
794, 3
Vokrouhlicky, D. 1998, A&A, 335, 1093
Vokrouhlick´y, D. 1999, A&A, 344, 362
Vokrouhlick´y, D., & Farinella, P. 1998, AJ, 116, 2032
Vokrouhlick´y, D., Chesley, S. R., & Milani, A. 2001, Celestial
Mechanics and Dynamical Astronomy, 81, 149
Voyatzis, G., Hadjidemetriou, J. D., Veras, D., & Varvoglis, H.
2013, MNRAS, 430, 3383
Wilson, D. J., Gansicke, B. T., Koester, D., et al. 2014, MN-
RAS, 445, 1878
Wyatt, S. P., & Whipple, F. L. 1950, ApJ, 111, 134
Wyatt, M. C., Farihi, J., Pringle, J. E., & Bonsor, A. 2014,
MNRAS, 439, 3371
Zuckerman, B., & Becklin, E. E. 1987, Nature, 330, 138
Zuckerman, B., Koester, D., Melis, C., Hansen, B. M., & Jura,
M. 2007, ApJ, 671, 872
c(cid:13) 2015 RAS, MNRAS 000, 1 -- 8
|
1602.03037 | 1 | 1602 | 2016-02-09T15:44:29 | Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks | [
"astro-ph.EP",
"physics.flu-dyn"
] | Planetesimals in gaseous protoplanetary disks may grow by collecting dust particles. Hydrodynamical studies show that small particles generally avoid collisions with the planetesimals because they are entrained by the flow around them. This occurs when $St$, the Stokes number, defined as the ratio of the dust stopping time to the planetesimal crossing time, becomes much smaller than unity. However, these studies have been limited to the laminar case, whereas these disks are believed to be turbulent. We want to estimate the influence of gas turbulence on the dust-planetesimal collision rate and on the impact speeds. We used three-dimensional direct numerical simulations of a fixed sphere (planetesimal) facing a laminar and turbulent flow seeded with small inertial particles (dust) subject to a Stokes drag. A no-slip boundary condition on the planetesimal surface is modeled via a penalty method. We find that turbulence can significantly increase the collision rate of dust particles with planetesimals. For a high turbulence case (when the amplitude of turbulent fluctuations is similar to the headwind velocity), we find that the collision probability remains equal to the geometrical rate or even higher for $St\geq 0.1$, i.e., for dust sizes an order of magnitude smaller than in the laminar case. We derive expressions to calculate impact probabilities as a function of dust and planetesimal size and turbulent intensity. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. turbulence_and_dust
February 10, 2016
c(cid:13)ESO 2016
Effect of turbulence on collisions of dust particles with
planetesimals in protoplanetary disks
H. Homann1, T. Guillot1, J. Bec1, C. W. Ormel2, S. Ida3, 4, and P. Tanga1
1 Laboratoire J.-L. Lagrange, Université Côte d'Azur, Observatoire de la Côte d'Azur, CNRS, F-06304 Nice, France
e-mail: [email protected]
2 Anton Pannekoek Institute for Astronomy, University of Amsterdam, The Netherlands
3 Department of Earth and Planetary Sciences, Tokyo Institute of Technology, 152-8551 Tokyo, Japan
4 Earth-Life Science Institute, Tokyo Institute of Technology, 152-8550 Tokyo, Japan
6
1
0
2
b
e
F
9
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
7
3
0
3
0
.
2
0
6
1
:
v
i
X
r
a
DRAFT: Do not distribute, February 10, 2016
ABSTRACT
Context. Planetesimals in gaseous protoplanetary disks may grow by collecting dust particles. Hydrodynamical studies show that
small particles generally avoid collisions with the planetesimals because they are entrained by the flow around them. This occurs
when St, the Stokes number, defined as the ratio of the dust stopping time to the planetesimal crossing time, becomes much smaller
than unity. However, these studies have been limited to the laminar case, whereas these disks are believed to be turbulent.
Aims. We want to estimate the influence of gas turbulence on the dust-planetesimal collision rate and on the impact speeds.
Methods. We used three-dimensional direct numerical simulations of a fixed sphere (planetesimal) facing a laminar and turbulent
flow seeded with small inertial particles (dust) subject to a Stokes drag. A no-slip boundary condition on the planetesimal surface is
modeled via a penalty method.
Results. We find that turbulence can significantly increase the collision rate of dust particles with planetesimals. For a high turbulence
case (when the amplitude of turbulent fluctuations is similar to the headwind velocity), we find that the collision probability remains
equal to the geometrical rate or even higher for St ∼> 0.1, i.e., for dust sizes an order of magnitude smaller than in the laminar case.
We derive expressions to calculate impact probabilities as a function of dust and planetesimal size and turbulent intensity.
Key words.
formation, Planet-disk interactions, Turbulence, Accretion, Accretion disks, hydrodynamics,
methods: numerical
Planets and satellites:
1. Introduction
Conventional models for planet formation involve the hierarchi-
cal growth by accretion of "planetesimals". Such building blocks
undergo collisions and gravitational binding to eventually reach
planetary sizes (see, e.g., Safronov 1972; Hayashi et al. 1985).
Still several serious uncertainties remain in the processes leading
from sub-µm dust grains, which follow the sub-Keplerian gas,
to km-sized planetesimals that are massive enough to move on
Keplerian orbits. One of them is known as the "meter-size bar-
rier". Meter-sized bodies experience a strong drag force, which
causes rapid orbital decay due to angular momentum loss. The
timescale of this decay is less than 100 years at 1AU (see, e.g.,
Weidenschilling 1984; Nakagawa et al. 1986), which is shorter
by several orders of magnitude than the observationally inferred
disk lifetime (several million years). Consequently, the planet-
forming material are lost from the disk.
One scenario for overcoming the meter-size barrier is by
the combined effect of streaming instabilities and pebble accre-
tion. Because of growth to mm/cm-sized particles and settling,
dust concentrates in the midplane and may become unstable
to streaming instabilities (Youdin & Goodman 2005; Johansen
et al. 2007; Johansen et al. 2011). This forms relatively large
planetesimals. After that, these large planetesimals can sweep
up surrounding grains and migrating pebbles (Ormel & Klahr
2010; Ormel & Kobayashi 2012; Lambrechts & Johansen 2012).
Because the orbital decay of the pebbles is also fast, a large num-
ber of pebbles are supplied from outer disk regions on relatively
short timescales, resulting in rapid planet growth. This scenario
utilizes the too rapid migration problem, rather than avoids it.
The pebble accretion model is actively discussed for the forma-
tion of the solar system (Morbidelli et al. 2015) and of close-in
super-Earth systems in exoplanetary systems (Chatterjee & Tan
2014, 2015). However, the details of the the accretion of grains
and pebbles by (large) planetesimals have not been fully clari-
fied.
Guillot et al. (2014) proposed detaile expressions for the ac-
cretion rates of dust grains in a protoplanetary gaseous disk for
a wide range of dust and planetesimal sizes. They point out that,
when using the numerical results by Sekiya & Takeda (2003),
the accretion probability drops off by several orders of magni-
tude at relatively small dust grains (within what they refer to
as the "hydrodynamical regime"). This strong depletion can be
easily explained by qualitative physical arguments. The mo-
tion of small grains is strongly coupled to the gas flow. Since
they follow gas streamlines they may avoid (head-on) collisions
with the planetesimal. However, the numerical simulations by
Sekiya & Takeda (2003) assume laminar flow, while it is known
that the disk gas is most likely to be turbulent. Observation-
ally inferred accretion rate in T-Tauri disks is far higher than
that caused by molecular viscosity. Consequently, protoplane-
tary disks are expected to be turbulent, with a turbulent eddy
viscosity νt much higher than the molecular kinematic viscosity.
Article number, page 1 of 15
Turbulence should affect the reduction in the dust accretion prob-
ability in hydrodynamical regime. Recently, Mitra et al. (2013)
studied the influence of turbulence on the dust impact velocity by
means of two-dimensional direct numerical simulations. They
found a velocity distribution with exponential tails and argued
that most of the dust collides with a speed comparable to that of
the head wind.
The purpose of this paper is to propose expressions for dust
accretion probabilities on planetesimals in turbulent gas, based
on three-dimensional direct hydrodynamical simulations. Grav-
ity effects are omitted in this study and are left for a future work.
We follow the approach introduced by Homann & Bec (2015)
and consider an idealized situation in which the planetesimal
is assumed to be spherical with a smooth surface. The gas is
modeled as a purely hydrodynamical and incompressible fluid so
that other possible modes originating from either the magneto-
rotational instability (MRI) or compressibility are not taken into
account. Additionally, we neglect any shear, disregard gravi-
tational effects, and do not allow for rotation of the planetesi-
mal. The aim of this work is twofold: First, it shall provide a
detailed study of collisions between small inertial grains and a
large spherical object and, secondly, it shall determine its conse-
quences on the accretion of dust by planetesimals in the context
of planet formation.
It is worth mentioning that our results, involving the hydro-
dynamical interactions and the collisions between a large spheri-
cal inclusion and small particles with inertia, have important ap-
plications in contexts that go beyond planet formation. In atmo-
spheric physics, this problem is known as "inertial impaction"
and is relevant for estimating rates in rain formation or wet de-
position of aerosols where a falling water drop scavenges smaller
cloud droplets or solid pollutants. Studies of such problems of-
ten rely on collision efficiencies and, again, mainly formulas
from laminar flow conditions are used (Berthet et al. 2010). A
popular formula is that of Slinn (1974), who proposed a fit of the
collision efficiency for the inertial impaction regime. Later, he
included molecular diffusion and extended his formula to smaller
projectiles (Slinn 1976, 1983). Inertial impaction is also impor-
tant for the design and improvement of industrial filters. Haugen
& Kragset (2010) studied the impact of particles on a cylinder
in a two-dimensional laminar inflow. Later, Hydle Rivedal et al.
(2011) investigated the case of a turbulent inflow and found that
turbulent fluctuations yield up to ten times more collisions than
a corresponding laminar inflow.
The paper is organized as follows. In Section 2, we explain
the disk model and notations. In Section 3, we describe the nu-
merical method that we use in our approach. In Section 4, we
present the results of our hydrodynamical simulations. In par-
ticular, we propose an expression for the accretion probability
in laminar and turbulent flows as a function of both the Stokes
number, which compares dust inertia to the perturbation of the
gas motion due to the planetesimal, and the turbulent intensity of
the surrounding gas flow. Section 5 is devoted to astrophysical
discussions on the obtained results. Section 6 contains a sum-
mary and some concluding remarks.
2. Protoplanetary disk turbulence model &
notations
2.1. Disk model
We consider a minimum mass solar nebula (MMSN) model
(Weidenschilling 1977b; Hayashi 1981), where the surface den-
sity and temperature of the gas in the disk are given in terms of
Article number, page 2 of 15
power laws:
Σgas = Σ1
Tgas = T1
(cid:18) r
(cid:18) r
au
au
(cid:19)−3/2
(cid:19)−1/2
,
.
(1)
(2)
Σ1 and T1 are the values at 1 a.u., for which we choose Σ1 =
1700 g cm−2 and T1 = 270 K. (A list of symbols and their def-
initions used in this paper is summarized in the appendix A.)
Assuming the disk is vertically isothermal, with the isothermal
kBTgas/µ ∝ r−1/4 and a mean molecular
sound speed cs = (cid:112)
(cid:34)
weight of µ = 2.34mH, the density reads
(cid:19)2(cid:35)
ρ(r, z) =
Σgas(r)
H √2π
exp
−
1
2
(cid:18) z
H
,
(3)
(cid:112)
GM(cid:63)/r3
where H = cs/ΩK is the disk scale height and ΩK ≡
the orbital (Keplerian) frequency at distance r from the star. It
follows that H/r ∝ r1/4, so that the disk is flared.
Due to the density and temperature gradients, the disk is
slightly supported by pressure. The amount of pressure support
(cid:33)2
∆g = (dP/dr)/ρ compared to the solar gravity is often expressed
in terms of a parameter η defined as
η ≡ −
∆g
2g(cid:63)
= −
= 1
2
= 1
2
d log P
d log r
P
2ρ Ω2
K r2
h2 ∇logP ≈ 1.63 h2,
∇logP
(4)
(cid:32) cs
vK
where we introduced the pressure logarithmic gradient ∇logP ≡
−d(log P)/d(log r), the Keplerian velocity vK ≡ ΩK r, and the
disk aspect ratio h ≡ cs/vK = H/r. The numbers are obtained
from an MMSN disk with the above power law profiles for den-
sity and temperature. It follows that the motion of the disk is less
than Keplerian (Weidenschilling 1977a); the gas flows at a speed
equal to (1−η) vK. Also, the headwind that is faced by a big body
(a.k.a. planetesimal), which moves at a Keplerian speed through
the sub-Keplerian rotating gas, then reads
Uc = η vK = 52 m s−1
(5)
which, for an MMSN disk is independent of the distance r to the
star.
However,
2.2. Turbulence model
The gas has a mean-free path of (cid:96)mfp = µ/(√2 σmol ρ) where
σmol ≈ 2 × 10−15 cm2 (Chapman & Cowling 1970) is the molec-
ular cross section. The kinematic molecular viscosity of the gas
is νmol = (1/2) vth (cid:96)mfp where vth = √8/π cs is the mean molecu-
lar thermal speed.
these parameters the resulting diffusion
for
timescales ∼ r2/νmol are too long to explain, for example, the
measured accretion rate in T-Tauri disks. Consequently, pro-
toplanetary disks are expected to be mildly turbulent, but still
subsonic, with a turbulent eddy viscosity νt assumed to be much
larger than the molecular kinematic viscosity. An often used pa-
rameterization is the alpha-viscosity model of Shakura & Sun-
yaev (1973):
νt ≈ α cs H.
(6)
The dimensionless constant α is a disk-dependent parameter. Its
value may range from a minimum of α ≈ 10−5, when the turbu-
lence originates from the Kelvin-Helmholtz instability of a dust
H. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
(cid:18) r
(cid:19)−3/2
1 au
layer at the mid-plane, to perhaps α ≈ 0.1 for the most violent
disks with a strong magneto-rotational instability (the MRI; Bal-
bus & Hawley 1991). Generally, the saturation level of the MRI
turbulence depends on the magnitude of the magnetic field that
threads the disks, the level of ionization, and the amount of dust
(Turner et al. 2014). Consequently, a variety of α-values can be
expected (Ormel & Okuzumi 2013). But even for a weak level
of turbulence – meaning, α (cid:28) 1 – the Reynolds number of the
flow, Re = νt/νmol, is very large by virtue of the very low densi-
ties that characterize astrophysical environments. Indeed, for an
MMSN density profile:
Re = 2 α
H
(cid:96)mfp ≈ 2.4 × 1012 α
.
(7)
For such expected large values of the Reynolds number,
one can reasonably assume that the gas flow is in a developed
three-dimensional turbulent regime which can be described us-
ing Kolmogorov (1941) phenomenology. The kinetic energy is
injected at a large (integral) scale L by a hydro- or magneto-
hydrodynamical instability. In both cases the injection mecha-
nism is associated to the sub-Keplerian shear. The associated
shear rate sets the typical turbulent large-eddy turnover time to
tL ∝ Ω−1
K (Cuzzi et al. 2001). The eddy viscosity then reads
νt = L2/tL = L2 ΩK leading, together with Equation (6), to
L = α1/2 H and vL ≡ L/tL = α1/2 cs. Note that the assump-
tions of incompressibility (vL (cid:28) cs) and three-dimensionality
(L (cid:28) H) of the turbulent flow in the disk are both fulfilled as
long as α (cid:28) 1. According to Kolmogorov (1941) phenomenol-
ogy, the kinetic energy cascades downscale with a constant rate
εKol = v2
ΩK until it reaches the smallest active scales,
the Kolmogorov scale (cid:96)Kol below which it is dissipated by molec-
ular viscosity. The typical velocity v(cid:96) of eddies whose size (cid:96)
lies in the inertial range (cid:96)Kol (cid:28) (cid:96) (cid:28) L reads v(cid:96) (cid:39) (εKol (cid:96))1/3.
The Kolmogorov length is defined as the scale where the scale-
dependent Reynolds number Re((cid:96)) ≡ v(cid:96) (cid:96)/νmol is unity, so that
(cid:96)Kol = ν3/4
mol/ε1/2
Kol.
The integral Reynolds number Re = νt/νmol gives the extension
of the inertial range: L/(cid:96)Kol = Re3/4 and tL/tKol = Re1/2.
Kol and the associated timescale tKol = ν1/2
L/tL = α c2
s
mol/ε1/4
2.3. Dimensionless quantities
Let us now turn back to our original problem,
that is the
aerodynamic interactions between the disk gas flow and a
solid planetesimal of size Rp, diameter d = 2Rp. There are
three dimensionless quantities characterizing the system: (the
numerical values below are those obtained for an MMSN disk)
– The planetesimal Reynolds number
Rep ≡
dUc
νmol
= ∇logP h
√8/π
≈ 2.7 × 104
(cid:32) d
(cid:33)
(cid:96)mfp
(cid:32) d
(cid:33)(cid:18) r
km
1 au
(cid:19)−2.5
,
(8)
which corresponds to the strength of inertia with respect to
molecular viscous forces for the gas flow surrounding the
planetesimal and characterizes how turbulent is its wake.
– The turbulent intensity
vL
Uc
I ≡
=
α1/2 cs
(1/2) h2∇PvK
=
√4α1/2
h∇logP ≈ 20 α1/2
(cid:18) r
(cid:19)−1/4
1 au
,
(9)
Fig. 1. Turbulent disk parameters: Planetesimal Reynolds number Rep
(black), dimensionless turbulent intensity I (red), and planetesimal to
Kolmogorov scale d/(cid:96)Kol (blue) for a MMSN disk model, disk turbulent
parameters α = 10−4 (dashed) and α = 10−2 (solid) and planetesimal
diameter d = 1 km. The blue and black line thus scale up and down with
the planetesimal size. The shaded regions are covered by the numerical
experiments (see Table 1).
which measures how strong the turbulent velocity fluctua-
tions are compared to the speed of the planetesimal slip.
– the ratio between the planetesimal size and the Kolmogorov
scale
δKol ≡
d
(cid:96)Kol
(cid:33)(cid:32) H
(cid:33)3/4
(cid:32) d
(cid:19)−19/8
H
(cid:96)mfp
(10)
d
=
α1/2H
≈ 400 α1/4
Re3/4 = 23/4α1/4
(cid:32) d
(cid:33)(cid:18) r
km
1 au
which measures the range of turbulent eddies that might in-
terfere with the planetesimal perturbation of the gas flow. It
is indeed known that the flow perturbation occurs on scales
of the order of d, so that all turbulent eddies of sizes between
(cid:96)Kol and d can potentially modify the gas flow around the
spherical planetesimal.
Figure 1 shows the dependence of the various parameters
with the orbital distance for two values of α compared to the as-
sumptions used for the numerical calculations. We can see that
Rep decreases with increasing orbital distance (kinematic viscos-
ity goes up). Similarly, d/(cid:96)Kol decreases because (cid:96)Kol increases
(turbulent scales are getting larger). The turbulent intensity is
not a very strong function of disk radius, but it depends quite
strongly on the turbulence parameter α.
In the laminar case, Rep = 400 corresponds to a 1 km plan-
etesimal located at ∼ 5 au in the MMSN disk, or e.g. to a 10 km
planetesimal at ∼ 12 au.
For the turbulent case, ideally we should find parameters for
which all three lines match with parameters of the numerical ex-
periments. However, because of the ∼ 106 mismatch between
the Reynolds number in the disk and the maximum one that can
be attained by present-day numerical simulations, this is not yet
possible. We come back to this issue when applying the results
of numerical models to real disk conditions.
Article number, page 3 of 15
10−1100101102Diskradius[AU]10−210−1100101102103104Repd/lKolI3. Numerical model
3.1. Gas flow
We focus on the collision rate of a stream of dust particles with
one spherical planetesimal. To study this situation we perform
3D direct numerical simulations (DNS) of a hydrodynamic flow
around a spherical object with a no-slip boundary conditions at
its surface. The overall method consists in a combination of a
standard pseudo-Fourier-spectral solver with a penalty method
that is explained in this section.
We integrate the incompressible Navier-Stokes equations
(12)
(11)
1
ρg∇p + νmol∇2u + f , ∇ · u = 0,
∂tu = −u · ∇u −
for the gas velocity u, where ρg is the gas density, νmol its kine-
matic molecular viscosity and f a force. The latter maintains a
uniform inflow speed and an eventually ambient turbulent flow.
Its form is described later. The pseudo-Fourier-spectral approach
consists in computing spatial derivatives in Fourier-space and
convolutions arising from the non-linear terms in real space. A
Fast-Fourier transform is used to switch between the two spaces.
We use the P3DFFT library (see Pekurovsky 2012) that is very
efficient on massive parallel computers.
The Navier–Stokes equation is integrated in the reference
frame of the planetesimal. The spatial average of the velocity
field is thus fixed and given by the planetesimal speed Uc relative
to the gas. Equation (11) is associated with a no-slip boundary
condition at the surface of the planetesimal, which is assimilated
to a spherical object at rest whose diameter is denoted by d and
its center by XS . We thus have
u(x, t) = 0 for x − XS = d/2.
Numerically, this no-slip condition is enforced by an immersed
boundary technique (IBM). The latter technique was first used
to simulate the blood in flow in the context of a human heart by
Peskin (1972). Today, a variety of different approaches exists
(Mittal & Iaccarino 2005). Generally speaking, IBM consists in
solving fluid equations in domains with complex boundary con-
ditions such as moving heart valves on Cartesian grids. As such
boundaries are generally not grid conform, their effect on the
flow has to be modeled. For our problem of a spherical obstacle
at rest, the idea consists in defining the velocity field in the full
domain enforcing a vanishing velocity in the entire object, that
is for all x such that x − XS ≤ d/2. The Navier-Stokes equa-
tion (11) then has to be modified by introducing in its right-hand
side a penalty force fb(x, t), which acts as a Lagrange multiplier
associated to the constraint defined by the boundary condition
(12). The full problem (11)-(12) can then be rewritten as
∂tu = L(u) + fb, ∇ · u = 0,
where L(u) denotes the right hand side of (11). In order to com-
pute fb we make use of a direct forcing method introduced by
Fadlun et al. (2000) where we directly impose the planetesimal
velocity to the grid using the technique of a pressure predictor
and an improved modeling of the spherical inclusion. Bench-
marks (see Homann et al. 2013) of this method for a fixed
sphere show good agreement with existing data. This method
has also been used for the study of moving neutrally buoyant
particles in homogeneous isotropic turbulence in Homann & Bec
(2010) and Cisse et al. (2013). A similar IBM has been used
by Uhlmann (2005) together with second order finite-difference
Navier-Stokes solver to simulate turbulent suspension involving
many particles.
(13)
Article number, page 4 of 15
3.2. Dust particles
Dust particles are modeled by spherical inertial particles with a
radius a much smaller than the smallest scales of the flow, so that
they can be approximated by point particles. Further, we assume
that these particles move sufficiently slow with respect to the
gas and that their mass density ρp is much higher than the gas
density ρg. With these assumptions the dominant hydrodynamic
force exerted by the gas is a drag force, which is proportional to
the velocity difference between the particle and the gas flow (see
Maxey & Riley 1983; Gatignol 1983)
X = 1
ts
(cid:104)
u(X, t)− X
(14)
(cid:105)
where the dots stand for time derivatives. ts is called the response
(or stopping) time and is a measure of particle inertia. It is the re-
laxation time of the particle velocity to that of the gas (see Guil-
lot et al. 2014, for a description of how the particle size relates to
the stopping time). Finally, we also assume that the particles are
sufficiently diluted to neglect any interaction among them and
any back-reaction on the flow.
Usually, particle inertia is quantified in terms of the Stokes
number St = ts / tc defined by non-dimensionalizing their re-
sponse time by a characteristic time scale tc of the carrier flow.
The present problem involves different relevant time scales. Td =
d/Uc, the time it takes a dust particle to pass the planetesimal,
tL, the turbulent large-eddy turn-over time and tKol, the turbu-
lent dissipation time scale. If not otherwise specified, we use
tc = Td as this time scale rules the collision efficiency in the
laminar regime: Particles with small St are closely coupled to
the flow and are swept around the obstacle. Large St particles
preferentially collide with it.1
In studies concerned with the small-scale dynamics of iner-
tial particles one usually uses tc = tKol (Bec et al. 2006) and we
denote the associated Stokes number StKol.
3.3. Simulation setup
We performed simulations of a planetesimal moving through a
uniform and different turbulent flows. The physical situation is
illustrated in Fig. 2 where the planetesimal, together with dust
particles are shown in a small slice. These are snapshots taken
from our three-dimensional simulations. The planetesimal ex-
periences a headwind speed Uc from the right that is advecting
the dust particles. The upper panel is taken from a simulation
with a uniform inflow while the two others include turbulence
with two different intensities. Dust particles that have collided
with the planetesimal are missing downstream, resulting in an
empty region behind the planetesimal that is clearly visible for
the laminar flow, but much less in the turbulent cases.
These figures show important differences between laminar
and turbulent gas flows. The spatial distribution of the dust as
well as its local velocity strongly depend on the carrier flow
type. Note that while St = 0.8 for all cases shown in Fig. 2,
StKol = 1.25 for I = 0.29 and StKol = 9.8 for I = 1.18 so that the
clustering properties of the particles change from one value of I
to the other.
The runs involving a turbulent flow are set up in the follow-
ing way (see table 2 for parameters and definitions). An initially
1 Note that in the context of astrophysical disks, a different Stokes
number is often defined from τ = tsΩK. The definition of St that is
used here is denoted by τf in e.g., Sekiya & Takeda (2003) and Guillot
et al. (2014).
H. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
Fig. 2.
Instantaneous streamlines of
the gas flow (entering from the right with
a speed Uc) around a planetesimal with
Rep = 400 and dust particles (St = 0.8)
for I = 0 (laminar, δKol = 0), I = 0.29
(moderately turbulent, δKol = 25), and
I = 1.18 (strongly turbulent, , δKol = 70)
from top to bottom.
smooth large scale flow is integrated according to (11) with-
out any forcing f. Once a turbulent flow has developed (after
approx. 1− 2 tL) the velocity field is forced by keeping con-
stant the energy content of the two lowest wave number shells
(1 ≤ k ≤ 2) in Fourier space. This leads to a statistically sta-
tionary turbulent flow to which the planetesimal is added. For
this, the root-mean-square value vL of the velocity fluctuations
is normalized to the values given in table 2 for each simulation.
A mean velocity of Uc in one direction is imposed by keeping
constant the zero Fourier mode of the corresponding component
of the velocity. The planetesimal is modeled via the mentioned
immersed boundary method. The integration is continued until a
statistically stationary state is reached again. At this point, iner-
tial particles are seeded at a constant rate into the flow in a plane
sufficiently far from the planetesimal so that they have enough
time (> ts) to relax to the flow. During the simulation we re-
move and record all the particles that are touching the spherical
planetesimal or reaching the end of the computational domain.
On average the domain is filled with approximately ten million
particles.
The laminar flow simulations (see table 1 for parameters and
definitions) start with a uniform flow (Uc = 1) in which the
planetesimal is placed. Disturbances produced by its wake are
removed at the end of the computational domain via another ap-
plication of the penalty method, so that they are not re-injected
upstream the planetesimal by the periodic boundary conditions.
The dust particles are introduced into the flow once a (statisti-
cally) stationary state is reached.
boundary layer, all turbulent scales in its wake and the smallest
scales of the possibly turbulent ambient flow.
All physical flow parameters are determined by three di-
mensionless parameters: the Reynolds number of the planetes-
imal Rep = Uc d/νmol, the Reynolds number of the gas Re =
vL L/νmol, L being the integral scale of the ambient turbulent
flow, and the turbulent large-scale intensity I = vL/Uc. The
latter measuring the strength of the large-scale turbulent fluctu-
ations compared to the mean flow velocity. In the laminar case
(Re = 0, I = 0), we varied the planetesimal Reynolds number
Rep from 100 to 1000. We analyze the effect of turbulent fluctu-
ations for one specific choice of the planetesimal Reynolds num-
ber, namely Rep = 400 (turbulent wake) and vary I from 0.14 to
1.18. The corresponding flow Reynolds numbers Re (listed in
Table 2) are a consequence of our particular choice of the ex-
ternal force. Freezing the energy content of the lowest shells in
spectral space does not allow for changing L but only vL that
enters in the definitions of both I and Re.
The particle dynamics is characterized by the Stokes number
St. In all simulations we consider streams of heavy dust particles
with response times ts in between 0.04 and 81.92, corresponding
to Stokes numbers St in the range 0.05 ≤ St ≤ 63. The main
parameters of all simulations are summarized in table 2 and ta-
ble 1.
4. Results
4.1. Laminar settings
The time integration of (11) uses a Runge-Kutta scheme of
third order. The grid resolution is chosen in order to resolve all
small scales of the problem: those of the spherical planetesimal
In a uniform gas flow all dust particles have the same velocity
far from the planetesimal. This physical situation is fully deter-
mined by the Reynolds number of the planetesimal Rep and the
Article number, page 5 of 15
Table 1. Parameters of the turbulence simulations. Rep = Uc d/νmol: planetesimal Reynolds number, Re = vLL/νmol: outer gas flow Reynolds
number, vL: root-mean-square velocity, εKol: mean kinetic energy dissipation rate, νmol: kinematic viscosity, d: planetesimal diameter, (cid:96)Kol =
(ν3
L/εKol: integral scale, tL = L/vL: large-
eddy turnover time, N3: number of collocation points.
mol/εKol)1/4: Kolmogorov dissipation length scale, tKol = (νmol/εKol)1/2: Kolmogorov time scale, L = v3
Rep
400
400
400
400
400
Re
200
450
600
900
3000
Uc
1
1
1
1
1
vL
0.14
0.29
0.39
0.60
1.18
εKol
9.2 · 10−4
7.7 · 10−3
1.95 · 10−2
6.9 · 10−2
4.8 · 10−1
νmol
0.002
0.002
0.002
0.002
0.002
d
0.8
0.8
0.8
0.8
0.8
(cid:96)Kol
0.054
0.032
0.025
0.018
0.0114
tKol
1.46
0.51
0.32
0.17
0.065
L
2.92
3.04
3.09
3.10
3.40
tL
21.0
10.6
7.9
5.2
2.9
Nx × Ny × Nz
Np
256 × 256 × 2048 ≈ 107
256 × 256 × 2048 ≈ 107
256 × 256 × 2048 ≈ 107
256 × 256 × 2048 ≈ 107
256 × 256 × 2048 ≈ 107
Table 2. Parameters of the laminar simulations. Rep = Uc d/νmol: plan-
etesimal Reynolds number , d: planetesimal diameter, νmol: kinematic
viscosity, Nx × Ny × Nz: number of collocation points; The following
parameters apply to all the three simulations: lengths of the computa-
tional domain lx × ly × lz = 2π× 2π× 16π, inflow speed Uc = 1, number
of small particles Np ≈ 107
Rep
100
400
1000
d
0.8
0.8
0.65
νmol
2 · 10−3
2 · 10−3
6.5 · 10−4
Nx × Ny × Nz
256 × 256 × 2048
256 × 256 × 2048
512 × 512 × 4096
those particles that moved around the planetesimal from getting
into the recirculation region of the wake. The few records ob-
served might be numerical noise.
Fig. 4.
Probability density function of the stream-wise position z at
which colliding particles impact the planetesimal. The center of the
planetesimal is located at z = 0 and separates its front (negative z) from
its back (positive z). The 'ballistic' curve refers to straight dust trajec-
tories that do not feel any hydrodynamic forces.
A central quantity of the accretion problem is the collision
efficiency (or collision rate) E(St) that is the ratio between the
actual collision rate and that expected for free-streaming parti-
cles that do not feel any hydrodynamic force from the gas flow.
The latter rate is obtained from the geometrical cross-section of
the planetesimal and reads nd Uc π d2/4, where nd designates the
number density of dust particles. We use the following notation:
E0 denotes the uniform gas flow case that we are going to ana-
lyze in this section and EI denotes the turbulent case parameter-
ized by the turbulent intensity I that is discussed in the following
section. The efficiency E0 is a monotonously increasing function
of the dust particle Stokes number St that asymptotically reaches
unity for particles with very large inertia (see Fig. 5). The higher
is Rep the more probable are collisions, especially at small St.
In this low inertia limit, E0 sharply drops for all Rep. The rea-
son for this is the well established existence of a critical Stokes
number Stc below which no collisions occur at all (Taylor 1940).
Indeed, in the case of an inviscid Euler flow (zero viscosity,
(cid:112)
Rep = ∞) it can be shown that E0 drops to zero for small but
finite StEuler
= 1/24 (Ingham et al. 1990). The viscous bound-
Rep
ary layer of a finite Rep planetesimal that shrinks as ∼ 1/
reduces the collision probability and increases Stc. For the lim-
iting case of a Stokes flow Michael & Norey (1970) computed
numerically a critical Stokes number of 0.605.
c
Fig. 3.
Projected frontal view of the average probability density of
impaction on the planetesimal for dust particles with St = 0.2. White:
Many collisions; Black: No collisions.
Stokes number St of the dust particles. A typical flow pattern is
illustrated in the upper panel of Fig. 2 where a planetesimal flies
through a stream of dust particles while creating a moderately
turbulent wake. If an approaching dust particle collides with the
planetesimal or not is only determined by its Stokes number and
impact parameter. Without the hydrodynamic flow that deflects
dust around the obstacle it would just be the impact parameter
ruling the collision rate so that the hydrodynamic forces reduce
the collision rate below that of the geometric cross section.
The colliding dust particles preferentially hit the planetesi-
mal on the axis of symmetry with a decreasing probability to
its edge. Small inertial particles only touch the planetesimal in
a central region leaving eventually an outer non-collisional ring
(see Fig. 3). But this region already disappears for Stokes num-
bers around unity, so that dust collisions fill the complete front
of the spherical planetesimal. In the limit of infinite S t the distri-
bution become uniform. Rear collisions virtually never happen
(see Fig. 4) in a laminar flow that is because inertia prevents
Article number, page 6 of 15
H. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
ESlinn
StSlinn
0
c
Slinn (1974, 1976, 1983) proposed a fitting function of the
collision rate of the form
(cid:32)
(cid:33)3/2
St − StSlinn
c
(St, Rep) =
(Rep) = 0.6 + (1/24) log(1 + Rep/2)
St − StSlinn
1 + log(1 + Rep/2)
+ 2/3
c
(15)
.
which showed a marginal improvement to the experimental and
numerical data available at the time. The uncertainties that we
obtain here are much smaller and allow to discriminate against
the fit from eq. (15).
c
StSlinn
is thus a critical number below which the (extremely low)
collision probability is determined by different physical mecha-
nisms. These expressions appear to have been fitted empirically
to both experimental and numerical data with uncertainties of
order 0.1 on the determination of E0(St) (see Slinn 1974).
Fig. 6.
Collision efficiency as a function of St − Stc, where Stc is
the critical Stokes number below which no collision happen. fit=(St −
Stc)/(St−Stc +2/3) (eq. (16)). The solid line indicates a linear increase.
Inset: Critical Stokes number as a function of the planetesimal Reynolds
number. fit=StSlinn
(Rep/3).
c
Fig. 5.
Collision efficiency as a function of the dust particle Stokes
number obtained from direct numerical simulations with uniform in-
flows (symbols). The lines correspond to the fitting formula (15) pro-
posed by Slinn (1983).
To study the large-Stokes number behavior in details it is
useful to analyze 1−E0(St) that measures how the free-streaming
particle limit is recovered as a function of St. All curves for
different Rep (see Fig. 7) fall on the top of each other and reveal
a St−1 behavior at large St. Again, we observe slight deviations
to Slinn's fitting formula, while our proposed expression (16)
fairly matches the data.
It is useful to look in more detail at the small and large St
efficiencies separately. In Fig. 6 the collision efficiency is shown
as a function of St − Stc and not simply as a function of St. We
hence focus on the behavior of the collision probability when ap-
proaching the critical Stokes number. Stc is chosen in such a way
that all curves fall on top of each other so that differences for dif-
ferent Reynolds numbers disappear. Our estimated Stc are close
to that of Slinn (15) (see inset of Fig. 6). The case of large but
finite values of Rep has also been addressed by Phillips & Kaye
(1999), using matched asymptotics together with numerical sim-
ulations of the particle dynamics inside the obstacle boundary
layer. They found critical Stokes numbers slightly larger than
those of Slinn (15), so that our value are in between these two
predictions.
E0 increases linearly at small St − Stc (see Fig. 6), that is in
contradiction to Slinn's formula (15) that predicts E0 ∼ (St −
Stc)3/2. To incorporate this small-Stokes behavior we propose
the fitting function
E0(St, Rep) =
St − Stc
Stc(Rep) = 0.6 + (1/24) log(1 + Rep/6)
St − Stc + 2/3
1 + log(1 + Rep/6)
which is in excellent agreement with our data. Here we mention
(as already remarked by Slinn) that an exponential fit of the form
exp(−a/St) also works quite well for not too small St but this
form does evidently not conform to a critical Stokes number.
We also note that Slinn (1974) had also proposed a similar
fit (linear instead of with a 3/2 exponent), but opted for eq. (15)
Fig. 7. Collision efficiency deficit 1−E0(St) as a function of St for dif-
ferent planetesimal Reynolds numbers and the various fitting formula,
as labeled.
(16)
4.2. Turbulent settings
One can expect that turbulent velocity fluctuations of the gas,
resulting in dust velocity fluctuations, alter the collision statistics
of dust and planetesimals. An analysis of these changes is the
subject of this section.
The turbulent accretion problem involves more parameters
than the laminar problem presented in the previous section. It
is determined by four dimensionless parameters. The planetes-
imal and dust properties are of course still described by the
Article number, page 7 of 15
10−310−210−11000.1110E0(St)StRep=100Rep=400Rep=1000Rep=100S i1983Rep=1000S i198310−310−210−11000.0010.010.1110100E0(St)St−Stc00.10.20.30.40.50.60.710100100010000StcRepRep=100Rep=400Rep=1000S i1983(cid:28)ex(cid:28)S i1983(cid:28)10−210−11000.11101001−E0(St)StRep=100Rep=400Rep=1000S i1983(cid:28)ex (cid:28)Reynolds number Rep and the dust Stokes number St, but turbu-
lence adds two additional dimensionless parameters specifying
the ambient turbulent flow. They are the gas Reynolds num-
ber Re = vL L/νmol and the turbulent intensity I = vL/Uc that
compares the amplitude of turbulent fluctuations with the mean
velocity of the gas.
In this work, we explicitly vary St and the turbulent inten-
sity I and fix the planetesimal Reynolds number to Rep = 400
(weakly turbulent wake) in order to limit computational costs.
The gas Reynolds number Re is implicitly varied as it is cou-
pled to I due to our specific forcing scheme that prescribes L to
approximately half of the domain size.
The physical situation for two different turbulent intensities
is illustrated in the mid and bottom panel of Fig. 2. Dust parti-
cles are in these cases advected by a chaotic and irregular flow
possessing coherent structures, i.e. structures eventually persist-
ing for a long time. This has two important implications: First,
the velocity of dust is fluctuating spatially and temporally (see
Fig. 2 b) and c)) and in turn (as we study in details below) mod-
ifies the collision statistics. Second, from studies of hydrody-
namic turbulence it is known that inertial particles tend to escape
from coherent rotating regions of the flow and tend to cluster in
straining regions (Squires & Eaton 1991). These agglomerations
are called preferential concentrationsShaw (2003); Balachandar
& Eaton (2010). However, we expect that these concentrations
play a negligible role for the present study in which we are con-
cerned with averaged accretion quantities such as the collision
efficiency. The reason is that the temporally averaged dust con-
centration in the ambient flow is approximately the same as in
the laminar case (not shown) so that particle density fluctuations
average out.
4.2.1. Collision efficiency
For the smallest Stokes numbers and the largest turbulent
intensity we observe more than one hundred times more colli-
sions than in the reference laminar flow. This relative increase
∆E(St, I) = EI/E0 − 1 of the turbulent efficiency compared with
the laminar one is shown in Fig. 9. It is represented as a function
of St− Stc that is the distance from the critical Stokes number Stc
as ∆E diverges when St → Stc. The large values of ∆E at small
St decreases as a power law with an exponent close to −1. The
dependence of ∆E on I is nearly quadratic. Indeed, the different
curves almost fall on the top of each other when normalized by
I1.7 as can be seen from the inset of Fig. 9.
Fig. 9.
Relative increase of the collision efficiency for several tur-
bulent intensities as a function of the distance from the critical Stokes
number of the laminar case.
Turbulent collision efficiencies (compare Fig. 8) are well fit-
ted by the function
EI(St) = exp(−a(I)/St) (1 + b(I) St/(1 + St2))
(17)
containing two parameters a and b. The dependence of these
parameters on the turbulent intensity is shown in Fig. 10 to-
gether with simple fitting formulas. The parameter a is decreas-
ing roughly exponentially as a function of I. b is close to zero for
small intensities and strongly increasing starting from I ≈ 0.4.
Fig. 8.
Collision efficiencies for several turbulent intensities I =
vL/Uc and Rep = 400. The I = 0 curve is the same as in Fig. 5. Solid
lines: fits of the function EI(St) = exp(−a(I)/St) (1 + b(I) St/(1 + St2))
with a = 0.49, b = −0.095 for I = 0.14; a = 0.38, b = −0.04 for
I = 0.29; a = 0.21, b = 0.31 for I = 0.60, a = 0.085, b = 1.08 for
I = 1.18.
Turbulent fluctuations significantly increase the collision
probability. Figure 8 shows the collision efficiency EI for vari-
ous I. One observes that higher is the turbulent intensity I, more
collisions happen. This increase is the strongest at small St and
disappears asymptotically at large St. As is discussed later, the
collision efficiency around St ≈ 1 remarkably exceeds unity.
Article number, page 8 of 15
Fig. 10. Dependence of the fitting parameter a and b on the turbulent
intensity I, for Rep = 400. fa = 0.65 exp(−1.8 I) is fitting function for
the parameter a, f l
= 1.4 (I − 0.4) (for I > 0.4) is a linear fit for the
b
parameter b. Both fits and the corresponding error bars were obtained
by a standard least square method.
In the small Stokes number limit, the turbulent collision effi-
ciency displays a clear exponential falloff (see Fig. 11) indicat-
ing that the critical Stokes number disappears when turbulence
10−210−11000.010.1110100EI(St)StI=0I=0.14I=0.29I=0.6I=1.18 0.200.20.40.60.811.200.20.40.60.811.2(cid:28)igaaeeabfaflbH. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
influences the dust motion. Additionally, curves for different I
fall approximately on top of each other once shown as a function
of StL = ts vL/L. It is thus a turbulent time scale, namely the
large-eddy turn-over time, that replaces the advection time d/Uc
relevant for the laminar problem.
effect can explain the observation of collision efficiencies ex-
ceeding unity in turbulent flows. For large St, particles move
ballistically so that they only sweep through the laminar (I = 0)
volume.
Fig. 11.
Logarithm of collision efficiencies showing the exponen-
tial character at small StL = ts vL/L. inset: Probability density function
(PDF) of stream-wise velocity gradient normalized to standard devia-
tion within the boundary layer.
There are different ways in which turbulence enhances the
collision probability. Fluctuations of the headwind speed mix
collision efficiencies for different I and St and allow for colli-
sions of dust particles beyond the critical Stokes number Stc. In
turbulent gas flows, the dust heads onto the planetesimal with
a fluctuating velocity. The probability distribution of the head-
wind is given by the one-point turbulent velocity distribution that
is known to be close to a Gaussian. For the present problem, its
standard deviation is given by the turbulent intensity I. These
headwind variations lead to fluctuating planetesimal Reynolds
number and dust Stokes numbers. Especially close to the lami-
nar value Stc this results in higher collision probabilities and an
absence of a critical Stokes number. Another consequence of
a variable headwind are fluctuations of the stream-wise veloc-
ity gradient σ in the upstream boundary layer of the planetesi-
mal. They modify the local Stokes number that can be defined by
St = ts σ. The probability for a collision of a dust particle with
stopping time ts is then P(St > Stc) = P(σ > Stc/ts), and thus
relates to the probability of observing a large velocity gradient
at the particle surface. The probability density function (PDF)
of σ shown in the inset of Fig. 11. The tails are close to expo-
nential, although we cannot rule out stretched exponential tails
as observed in homogeneous isotropic turbulence. Exponential
tails mean P(σ > Stc/ts) ∼ exp(−C Stc/ts) which is consistent
with the formally observed exponential fall off of EI(St) at small
values of St.
Another mechanism to increase the collision probability re-
lates to turbulent diffusion. This enhances the mobility of dust
and increases its flux onto the planetesimal. As a simplified
gedankenexperiment one can think of a sphere moving in a sinu-
soidal velocity field representing the large-scale turbulent fluc-
tuations. Let us assume (in the reference frame of the sphere)
a velocity of the form u = Ucex + I sin(2π/tL, t) ey. The vol-
ume swept by the sphere is evidently larger in the turbulent case
(I > 0) than in the laminar (I = 0). According to inertia, dust
particles are more or less coupled to the gas which makes the
swept volume additionally depending on St. Small St particles
stick to the gas trajectories so their swept volumes equal. This
Fig. 12.
position for I = 0.29 (top) and I = 0.6 (bottom).
Probability of collisions as a function of the stream-wise
We conclude this section with a study of the spatial distri-
bution of dust impacts on the surface of a planetesimal in a tur-
bulent disk. In Fig. 4 we saw that dust collisions preferentially
happen close to the stagnation point of the flow in a quiescent
disk. This is still true when turbulent fluctuation agitate the dust
(see Fig. 12). But the added randomness leads to a homogeniza-
tion of the impact position. The larger is I the more the collisions
fill the entire planetesimal surface. And especially for small St
particles, backward collision become frequent.
4.2.2. Impact velocity
The relative velocity of dust and a planetesimal at impact is cru-
cial for the dust accretion problem as it determines, together
with the angle of impact, the outcome of a collision. Low colli-
sion speeds lead to sticking of dust on the target surface, while
high speeds lead to bouncing, fragmentation with mass transfer
or erosion (e.g., Blum & Wurm 2008; Windmark, F. et al. 2012).
In laminar disks the mean impact speed of small-size dust
(with a stopping time smaller than the orbital period) is a
monotonously increasing function of inertia (Weidenschilling
1977a). Dust particles with inertia close to the critical Stokes
number only slightly touch the planetesimal surface while large
St particles collide with the full headwind speed. Dust particles
in turbulent disks experience gas velocity fluctuations and in turn
Article number, page 9 of 15
For a St = 0.4 dust particle in a I = 1.18 headwind, the velocity
distribution is up to six times broader than in a non-turbulent
disk.
Fig. 15.
Probability density function of the impact velocity normal
to the planetesimal surface for St = 0.8 and several turbulent intensities
as labeled. The bold lines correspond to measurements from numerical
simulations, while the thin lines refer to the non-central chi-squared
prediction (see text).
Fig. 16.
Probability density function of the surface normal impact
velocity for I = 0.29 and several Stokes numbers as labeled. As in
Fig. 15, the bold lines correspond to numerical simulations and the thin
lines refer to the non-central chi-squared prediction.
The impact speed has, besides its average value and stan-
dard deviation, a probability distribution that varies with both I
and St. It reveals that high-impact speeds of the order of several
times the headwind speed are quite probable. For a fixed value of
St the distribution becomes monotonously broader with increas-
ing I (see Fig. 15). The numerical data is there compared to the
non-central chi-squared distribution that would be obtained if vc
were the norm of a three-dimensional random Gaussian vector
with prescribed mean and variance. Up to statistical accuracy,
it seems from Fig. 15 that such an approach gives a rather good
description of actual fluctuations of the impact velocity. This
approximation is however valid only if the Stokes number St is
sufficiently large. Figure 16 indeed represents the same distribu-
tions for a fixed value of I at varying St. One observes devia-
tions from the chi-squared prediction in both tails at the smallest
value of the Stokes number. It seems nevertheless that the distri-
Fig. 13. Mean dust velocity at impact on the planetesimal surface for
Re = 400.
drag variations. They follow preferentially turbulent structures
with characteristic time-scales that equal their response time.
This coupling of inertial particles is known to create non-trivial
phenomena such as the mentioned small-scale preferential con-
centrations that are the most effective for StKol = ts/tKol ≈ 0.6
particles (Reade & Collins 2000; Bec et al. 2006).
We observe such a eddy-dust coupling also for the collision
velocity vc (the norm of the dust velocity vector at impact) of
dust particles with a planetesimal (see Fig. 13). Asymptotically,
small-St dust (small in the sense of particles with a small colli-
sion efficiency) still only mildly touches the surface, while large-
St dust collides with the speed of the headwind. However, at
intermediate values of St, turbulent velocity fluctuations lead to
an increase of the collision speed that even exceeds the head-
wind speed. For the highest turbulent intensity that is studied
here (I = 1.18), the average impact speed is approximately 75%
higher than the mean headwind speed. We remark that once
particle inertia is measured in terms of the characteristic time
scale of turbulent structures of size d (planetesimal diameter)
td = tL (d/L)2/3 (td = 4.4, 2.1, 1.1 for I = 0.29, 0.6, 1.18) all
maxima of the curves (located at S t ≈ 3.2, 1.6, 0.8) in Fig. 13
align at St = ts/td ≈ 0.6. Turbulent eddies of the size of the
planetesimal are thus responsible for this increase of the impact
velocity.
Fig. 14.
several turbulent intensities.
Standard deviation of the impact speed for Re = 400 and
To estimate the broadness of the velocity distributions we
measured in Fig. 14 their standard deviation. Here, the St ≈ 1
peak is even more important than for the mean impact speed.
Article number, page 10 of 15
v2c/hv2ci012345pdf10-310-210-1100I=0.14I=0.29I=0.60I=1.18v2c/hv2ci02468pdf10-310-210-1100St=0.10St=0.40St=12.80St=51.20H. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
bution still belongs to the same family and can be approximated
by a chi-squared distribution with a smaller number of degrees
of freedom. Everything happens as at small St, the dust velocity
fluctuations with respect to the planetesimal were constrained in
a space with dimension less than three.
The outcome of a collision also depends on the angle of im-
pact. Figure 17 shows the average collision angle ¯∠ with respect
to the surface normal direction n. Small St dust preferentially
touches the planetesimal with an mean impact angle close to 900.
High inertia dust heads straight onto the planetesimal and expe-
riences an average impact angle of 450. Turbulent fluctuations
randomize the impact angle and favor this way to the same mean
angle of ¯∠(v, n) = 450.
Fig. 18. Value of the factor fhydro (see eq. (18)) indicative of a reduction
of the linear cross section of planetesimals resulting from hydrodynam-
ical effects as a function of dust particle size, for a planetesimal of 1 km
radius in an MMSN disk, at an orbital distance of 7.1 au (plain) and
1 au (dashed) and for various levels of turbulence (as labeled). The re-
sults are compared to the one obtained by Guillot et al. (2014) which is
independent of turbulence amplitude. For the 7.1 au case (which corre-
sponds to Rep = 400 as in the previous hydrodynamical calculations),
the lines corresponding to α = 10−4 and 10−6 are hidden behind the
laminar case (see text).
Fig. 17. Average angle of impact ¯∠(v, n) with respect to the surface
normal direction n in degrees.
situation becomes similar to the laminar case, but with a head-
wind that is increased by √1 + I2. We write
(cid:105)1/2
(cid:104)
5. Astrophysical application
5.1. Linear cross section
fhydro =
E0(St∗, Rep)
[EI(St)]1/2
if δKol = d/(cid:96)Kol < 1
otherwise
(18)
We apply these results to realistic disk conditions. In order to
study filtering of dust by planetesimals, Guillot et al. (2014) had
applied the results of numerical simulations by Sekiya & Takeda
(2003) for a laminar disk and a fixed planetesimal Reynolds
number Rep = 50. Our simulations in the laminar case cover
a range of Rep from 100 to 1000. As seen in Fig. 1 and Eq. (8),
this corresponds to planetesimals with diameters between 4 m
and 40 m at 1 au and between 1 km and 12 km at 10 au. Since
the outcome of the simulations is only weakly dependent on Rep
(the critical Stokes number is a function of log(1 + Rep/2) –see
eq. (16)), we expect to be able to extrapolate the results outside
this range.
In most cases however, turbulence is expected to be impor-
tant. Our simulations in the turbulent case have been calculated
for various intensities of the turbulence I, but a fixed planetes-
imal Reynolds number Rep = 400. However, when turbulence
becomes important, we expect the results to become very weakly
dependent on Rep. This is for two reasons: First as seen in Fig. 2,
for values of I approaching unity, the flow around planetesimals
is perturbed very significantly and becomes controlled by the
turbulence of the disk instead of by the planetesimal properties.
Second, turbulence perturbs the boundary layer around the plan-
etesimal independently of its properties to offer new possibilities
for dust particles to impact.
But for small planetesimals and/or low turbulent intensity,
the planetesimal size can become smaller than the Kolmogorov
scale, i.e., the minimum scale for turbulent eddies. In that case,
the planetesimals experience a headwind of variable intensity
and direction. It is expected that, in the limit of d/(cid:96)Kol (cid:28) 1, the
p f 2
where fhydro is the collision efficiency as defined by Guillot et al.
(2014), St∗ = St √1 + I2and E0 and EI the fitting formula (16)
and (17) . With this definition, 2Rp fhydro is the planetesimal lin-
ear collisional cross section and πR2
hydro its surface collisional
cross section. Thus, for a planetesimal smaller than the smallest
turbulent eddy, the flow is considered laminar, but we account
with the use of St∗ for a flow velocity that is slightly higher on
average. On the other hand, when the planetesimal is larger than
the Kolmogorov scale, we use the results of the simulations in
the turbulent case directly.
Figure 18 illustrates how the factor fhydro varies as a func-
tion of particle size in an MMSN disk for a 1 km-radius plan-
etesimal either at 7.1 au or at 1 au. The first case corresponds
to a planetesimal Reynolds number Rep = 400 equal to the one
used in the hydrodynamical simulations with turbulence. The
second case corresponds to a much higher Reynolds number
Rep = 5.4 × 104 outside the range of our simulations.
In all cases, particles which are larger than a critical value
(i.e., with a Stokes number higher than unity) are accreted with
a cross-section approximately equal to the geometric one (i.e.,
fhydro ≈ 1). In laminar disks or when turbulence is small, the
cross section drops for small particles such that St < 1. If tur-
bulence is large enough, this drop occurs at even smaller sizes,
with an offset that corresponds to one to two orders of magnitude
for the case with α = 10−2.
The comparison of the laminar Rep = 400 cases shows a
relatively good agreement between our work and the previous
results of Guillot et al. (2014) who used results from Sekiya &
Takeda (2003) for a fixed planetesimal Reynolds number of 50.
Article number, page 11 of 15
10-710-610-510-410-310-210-1100Dust size [cm]10-610-410-2100fhydroRp=1 km, MMSNThis work, α=10−2This work, α=10−4This work, α=10−6This work, laminarGuillot et al. (2014)r=7.1 aur=1 auWhen turbulence is added, it is worth noticing that while the case
with α = 10−2 stands out and allows much smaller particles to
collide, the cases with α = 10−4 and 10−6 are almost indistin-
guishable from the laminar case. This is a direct consequence
from the fact that δKol < 1 for these: the smallest turbulent cell
is expected to be larger than the planetesimal size which implies
that we switch to the laminar case in eq. (18). Our approach is
thus discontinuous in α, but resolving this issue would require
dedicated simulations beyond the scope of the present work
At high Reynolds number (i.e., the 1 au case in Fig. 18), the
Kolmogorov parameter δKol is generally high which implies a
nearly continuous behavior from high to low values of α. A
small issue seen for low values of the viscosity and dust sizes
corresponding to Stokes number close to unity is that the value
of fhydro for a disk with low turbulence (e.g., α = 10−6) can
become smaller than the laminar value, which according to our
simulations is unlikely. Clearly, this is a consequence of the fact
that our expressions have been derived for a relatively low plan-
etesimal Reynolds number and are applied very far from that
value. Again, dedicated simulations would be needed, but may
be out of reach with present-day computing power.
5.2. Collision probabilities in disks
We now examine the consequences for collisions of dust grains
with planetesimals with the same approach as Guillot et al.
(2014). In protoplanetary disks, collisions between drifting dust
particles and planetesimals occur with a probability P3D that is a
function of the planetesimal cross section, the scale height of the
dust disk hd and of the drift velocity of the dust particles. The
latter depends on orbital distance r, orbital (keplerian) frequency
Ω and stopping time ts. In the limit that gravitational effects and
gas drift may be neglected and for circular orbits, this probability
can be shown to write (Guillot et al. 2014):
(cid:114)
1 + 1
4tsΩ
P3D = 1
2√π
p f 2
R2
hydro
Hdr
,
(19)
where fhydro accounts for hydrodynamical effects discussed pre-
viously (the purely geometrical limit is recovered for fhydro = 1).
In reality, gravity becomes important both for median to
large planetesimals (kilometer size and more) and for large
grains (above meter size) and seriously complicates the pic-
ture. Several interaction regimes may be defined as follows (see
Ormel & Klahr 2010; Guillot et al. 2014):
– The geometric regime, corresponds to the most simple case
in which drag, hydrodynamical and gravity effects may be
neglected.
– We define the hydrodynamical regime as an extension of this
regime at small dust sizes when we must account for the de-
flection of dust grains around planetesimals.
– The Safronov regime corresponds to the case when large
"dust" (effectively, boulders) which are very weakly affected
by gas drag migrate inward so slowly that they feel a grav-
itational focusing by the planetesimal which increases the
collision probability.
– In the three-body regime, the gravity fields of the planetesi-
mal and that of the central star must be taken into account.
– The settling regime corresponds to the case when gravita-
tional acceleration from the planetesimal and gas drag on the
dust particles lead to an enhanced capture probability.
Article number, page 12 of 15
Fig. 19. Contours of the collision probability P3D obtained at 1 au in
an MMSN disk with α = 10−2 as a function of dust size and planetesi-
mal size. The colored areas (labeled "geometric", "hydro", "Safronov",
"three-body" and "settling") correspond to different interaction regimes.
The top panel shows the results obtained using the same approach as
Guillot et al. (2014) but corrected for a factor fhydro (see text and com-
pare with their Fig. 16). The bottom panel corresponds to results with
the new prescriptions for the hydro model. For an easier comparison,
the dashed red line marks the location of the hydro regime (correspond-
ing to fhydro < 0.9) in the Guillot et al. (2014) study.
Accounting for the complexity of the problem we thus cal-
culate the collision probability between dust and planetesimals,
P3D, from eq. (43) of Guillot et al. (2014), assuming monodis-
perse size distributions2, but including fhydro from eq. (18). In
doing so, we also adopt an important modification stemming
from the work of Johansen et al. (2015) and Visser & Ormel
(2015): Instead of limiting the extent of the settling regime to
when the capture radius is larger than the physical size of the
planetesimal as in Guillot et al. (2014), we instead look for so-
lutions of the settling regime equations outside of this range and
adopt for the collision probability the maximum of the probabil-
ities obtained in the settling and geometric+hydro regimes. This
is important in regions where fhydro is extremely low but gravity
and gas drag can still affect the trajectories of the dust particles.
Figure 19 shows how P3D varies with dust and planetesimal
size for a fixed orbital distance of 1 au. We focus on planetesi-
mals smaller than 100 km and down to 10 m with the caveat that
2 In doing so, we correct for the fact that in Guillot et al. (2014), the
3D collision probability in the hydro mode was overestimated because it
neglected the reduction in the vertical cross section, i.e., P3D ∝ fhydroR2
p
had been assumed instead of P3D ∝ f 2
p. Because this affected
the hydro mode with an already very low collision probability this had
negligible effect on the qualitative results.
hydroR2
log(P3D) -1012Log(Planetesimal size) [km]Safronovsettling3-bodygeometrichydro-26-24-22-18-16-16-14-14-12-10-8-8Guillot et al. (2014)*-6-4-202Log(Dust size) [cm]-1012Log(Planetesimal size) [km]Safronovsettling3-bodygeometrichydro-20-18-16-16-14-14-12-12-10-10-8-8This work1.0 au, α=10-2, MMSN1.0 au, α=10-2, MMSNH. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
5.3. The inefficient capture of small dust grains
We now turn to the examination of how (in)efficiently individ-
ual small dust grains may have been collected by planetesimals
as a function of their sizes and orbital distance. A full model
would require studying the size distribution of dust and plan-
etesimals and is beyond the scope of the present paper. However,
we can identify the parameter space for which this collection of
dust is inefficient by identifying when the collision cross sec-
tion becomes smaller than 1% of the geometrical one (i.e., cor-
responding to fhydro < 0.1 in the limit when gravity effects are
not important). Because the drop in collision probability in the
hydro region of fig. 19 is quite abrupt, we expect that if dust is in-
deed collected individually by planetesimals, this process should
leave its imprint on the size distribution of individual grains in
meteorites.
Figure 20 identifies these regions as a function of orbital dis-
tance and planetesimal size, either in the case of a high turbu-
lence level (top panel) or a low turbulence level (bottom panel).
In both cases, the collection of very small particles (nanome-
ter sizes) is found to be very inefficient, at least inside of 10 au.
Dust particles of progressively larger sizes can be collected up to
shorter orbital distances, but the efficiency then strongly depends
on the turbulence level.
For 1 mm particles (corresponding to a typical size of chon-
drules), we do not see in fig. 20 a region of strongly inefficient
collection when turbulence is high (α = 10−2), but for α = 10−4,
these particles avoid collisions with planetesimals between about
0.3 and 30 km within a fraction of an au. One-micron particles
are collected inefficiently inside a region extending from about
0.1 to 3 au depending on planetesimal size for α = 10−2, but this
region extends from 0.8 to 6 au for α = 10−4. Smaller particles
can collide with planetesimals only at larger orbital distances,
when the gas density has decreased and the stopping time in-
creased for a given particle size.
A larger turbulence level can therefore allow collisions of
small-size particles which would otherwise be avoided due to the
hydrodynamical flow around the planetesimals. However, this is
also balanced by the fact that higher turbulence means a thicker
dust subdisk which lowers the collision probability (see Guillot
et al. 2014). Due to the form of eq. 18, the effect of turbulence
becomes weaker at large orbital distances, when the size of the
smallest turbulent eddies becomes larger than the planetesimals.
This thus explains why the contour lines for very small dust par-
ticles are identical for the α = 10−2 and α = 10−4 cases.
The change in behavior of the contour plots for α = 10−2,
dust sizes between 1 µm and 1 mm and orbital distances from
0.05 to 0.3 au is due to a change in drag behavior for these par-
ticles: At short orbital distances, the gas density is so high that
they are in the Stokes regimes and they switch to an Epstein drag
beyond about 1 au.
For the planetesimal sizes considered, particles smaller than
about 10 nm have a collision probability that is independent of
alpha. This is because collisions can occur only in the outer disk
where δKol < 1, i.e., the smallest Kolmogorov scale is still larger
than the planetesimals considered.
Small particles such as the presolar grains present in mete-
orites, which can have sizes of only a few nanometers (e.g., Clay-
ton & Nittler 2004) must have either collided with planetesimals
far out in the disk or be incorporated into larger grains which
would have themselves collided with planetesimals (e.g., Ormel
et al. 2008). For some of the presolar grains, given their very
low abundance, it remains possible that they were incorporated
directly in planetesimals, although this would have to be quan-
Article number, page 13 of 15
Fig. 20.
Regions in which the collection of dust grains becomes
very inefficient (defined as a collision cross section lower than 1% of
the geometric one –see text) as a function of orbital distance and plan-
etesimal radius for various dust sizes (labeled) and for two values of
the turbulence: α = 10−2 (top panel) and α = 10−4 (bottom panel).
Large particles are collected relatively efficiently everywhere except in
the innermost regions of the disk where the gas density is assumed to
be high. Small particles are inefficiently collected everywhere except at
the largest orbital distances, at least in the case of a disk characterized
by the MMSN gas density distribution.
for planetesimals smaller than about 1 km, gas drag should be
included. The top panel shows the previous results from Guillot
et al. (2014), which correspond to the case of a laminar flow and
no extension of the settling regime. The bottom panel shows the
results for a turbulent flow with full account for gravity effects
even for low-planetesimal sizes.
The comparison between the top and bottom panels of fig. 19
shows that even a weak planetesimal gravity effectively limits
the decrease of the collision probabilities in the extended set-
tling regime for dust smaller than ∼ 100 µm and planetesimals
between one and 100 km. The inclusion of turbulence effects
also shifts the hydrodynamic regime to smaller dust sizes. The
shift is about one order of magnitude for all planetesimal sizes
considered when comparing the results for α = 10−2 to those
for a laminar disk. Particles of 0.1 mm can hence be accreted
relatively efficiently by planetesimals for all the sizes consid-
ered. However, smaller particles still end in the hydrodynamical
regime with a strongly reduced collision efficiency. For exam-
ple micron-sized dust particles are very inefficiently captured by
planetesimals larger than a few kilometers in size.
-1012Log(Planetesimal size) [km] -1012Log(Planetesimal size) [km]> 100 µm> 10 µm> 1 µm> 100 nm> 10 nm> 1 nmMMSN, α=10−2−2−1012Log(orbital distance) [au]−1012Log(Planetesimal size) [km]−2−1012Log(orbital distance) [au]−1012Log(Planetesimal size) [km]> 1 mm> 100 µm> 10 µm> 1 µm> 100 nm> 10 nm> 1 nmMMSN, α=10−4tified. In any case, this should have occurred without leading
to any melting or dissociation of these grains which kept their
identify throughout.
6. Conclusions
We have derived the accretion probability of small particles by a
planetesimal in a turbulent gas. In order to do so, we performed
high-resolution hydrodynamical simulations of the flow around
a spherical planetesimal of diameter d moving with a velocity
Uc, assuming incompressibility. We studied both the case of
a laminar flow and that of a turbulent one, the intensity of the
turbulence being related to the turbulent viscosity of the disk.
Dust particles of variable size were implemented in the flow to
determine collision rates.
For laminar flows, we confirm that small particles with a
Stokes number St < 1 (corresponding to stopping times shorter
than the time to cross the planetesimal) see a significant drop in
their collision rate with the planetesimal. For turbulent flows
however, this drop occurs for sizes that can be significantly
smaller, i.e., turbulence helps accreting dust particles with sizes
up to one to two order of magnitudes smaller than for laminar
disks.
We thus derived collision probabilities both in the laminar
case [eq. (16)] and in the turbulent case [eq. (17)]. These ex-
pressions, even if limited to limited to Rep = 400, can be used
for a wide range of situations. We propose an approximate
recipe to use either the laminar case if the planetesimal size is
smaller than the Kolmogorov scale and the turbulent case other-
wise [eq. (18)].
When applied to real disks, our new expressions shift the
boundary with the hydro regime – where accretion rates are
greatly suppressed – to smaller sizes. For example, for α = 10−2,
the upper limit dust size in the hydrodynamical regime is de-
creased by a factor 100 and even sub-µm size particles collide ef-
ficiently with one-kilometer planetesimals. They also show that
the accretion of extremely small particles is difficult and gen-
erally requires to be done by small planetesimals (less than km
size) at large orbital distances (beyond 1 au) and/or late in time,
when the disk has become less massive. We believe that these re-
sults are important to interpret, among other things, the presence
and characteristics of presolar grains in meteorites since these
vary in size from several microns down to only a few nanome-
ters.
In order to apply our results to protoplanetary disks, we
had to approximate the effect of gravity, often by extrapolations
far from the regime in which numerical experiments were con-
ducted. Future efforts will be directed towards including the
gravitational force directly in our hydrodynamical simulations.
Appendix A: Definition of symbols
We summarize here the symbols used in this paper and their def-
initions.
Acknowledgements. We thank Satoshi Okuzumi, Zoe Leinhart and Rico Visser
for useful discussions. Most of the simulations were done using HPC resources
from GENCI-IDRIS (Grant i2011026174). Part of them were performed on the
"Mésocentre de calcul SIGAMM" at the Observatoire de la Côte d'Azur. The
research leading to these results has received funding from the Agence Nationale
de la Recherche (Programme Blanc ANR-12-BS09-011-04).
Article number, page 14 of 15
Berthet, S., Leriche, M., Pinty, J.-P., Cuesta, J., & Pigeon, G. 2010, Atmospheric
Research, 96, 325
Blum, J. & Wurm, G. 2008, ARA&A, 46, 21
Chapman, S. & Cowling, T. G. 1970, The mathematical theory of non-uniform
gases. an account of the kinetic theory of viscosity, thermal conduction and
diffusion in gases (Cambridge University Press)
Chatterjee, S. & Tan, J. C. 2014, The Astrophysical Journal, 780, 53
Chatterjee, S. & Tan, J. C. 2015, The Astrophysical Journal Letters, 798, L32
Cisse, M., Homann, H., & Bec, J. 2013, Journal of Fluid Mechanics, 735, R1
Clayton, D. D. & Nittler, L. R. 2004, ARA&A, 42, 39
Cuzzi, J. N., Hogan, R. C., Paque, J. M., & Dobrovolskis, A. R. 2001, ApJ, 546,
496
tional Physics, 60, 35
Fadlun, E. A., Verzicco, R., Orlandi, P., & Mohd, J. 2000, Journal of Computa-
Gatignol, R. 1983, J. Méc. Théor. Appl., 1, 143
Guillot, T., Ida, S., & Ormel, C. W. 2014, A&A, 572, A72
Haugen, N. E. L. & Kragset, S. 2010, Journ. Fluid Mech., 661, 239
Hayashi, C. 1981, Progress of Theoretical Physics Supplement, 70, 35
Hayashi, C., Nakazawa, K., & Nakagawa, Y. 1985, in Protostars and Planets
II, ed. D. Black & M. Matthews (University of Arizona Press, Tucson, AZ),
1100–1153
Homann, H. & Bec, J. 2010, J. Fluid Mech., 651, 81
Homann, H. & Bec, J. 2015, Physics of Fluids, 27, 053301
Homann, H., J.Bec, & Grauer, R. 2013, J. Fluid Mech, 721, 155
Hydle Rivedal, N., Granskogen Bjørnstad, A., & Haugen, N. E. L. 2011, ArXiv
e-prints
935
A62
Ingham, D., Hildyard, L., & Hildyard, M. 1990, Journal of Aerosol Science, 21,
Johansen, A., Klahr, H., & Henning, T. 2011, Astronomy & Astrophysics, 529,
Johansen, A., Mac Low, M.-M., Lacerda, P., & Bizzarro, M. 2015, Science Ad-
vances, 1, 15109
Johansen, A., Oishi, J. S., Mac Low, M.-M., et al. 2007, Nature, 448, 1022
Kolmogorov, A. 1941, Akademiia Nauk SSSR Doklady, 30, 301
Lambrechts, M. & Johansen, A. 2012, Astronomy & Astrophysics, 544, A32
Maxey, M. R. & Riley, J. 1983, Phys. Fluids, 26, 883
Michael, D. H. & Norey, P. W. 1970, Canadian Journal of Physics, 48, 1607
Mitra, D., Wettlaufer, J. S., & Brandenburg, A. 2013, The Astrophysical Journal,
Mittal, R. & Iaccarino, G. 2005, Annual Review of Fluid Mechanics, 37, 239
Morbidelli, A., Lambrechts, M., Jacobson, S., & Bitsch, B. 2015, Icarus, 258,
773, 120
418
Nakagawa, Y., Sekiya, M., & Hayashi, C. 1986, Icarus, 67, 375
Ormel, C. & Klahr, H. 2010, Astronomy & Astrophysics, 520, A43
Ormel, C. & Kobayashi, H. 2012, The Astrophysical Journal, 747, 115
Ormel, C. W., Cuzzi, J. N., & Tielens, A. G. G. M. 2008, ApJ, 679, 1588
Ormel, C. W. & Okuzumi, S. 2013, ApJ, 771, 44
Pekurovsky, D. 2012, SIAM Journal on Scientific Computing, 34, C192
Peskin, C. S. 1972, Journal of Computational Physics, 10, 252
Phillips, C. & Kaye, S. 1999, Journal of Aerosol Science, 30, 709
Reade, W. C. & Collins, L. R. 2000, Phys. Fluids, 12, 2530
Safronov, V. S. 1972, Evolution of the protoplanetary cloud and formation of the
earth and planets. (Israel Program for Scientific Translations, Keter Publish-
ing House)
Sekiya, M. & Takeda, H. 2003, Earth, Planets, and Space, 55, 263
Shakura, N. I. & Sunyaev, R. A. 1973, A&A, 24, 337
Shaw, R. A. 2003, Annual Review of Fluid Mechanics, 35, 183
Slinn, W. 1983, Atmospheric Sciences and Power Production, Chap.11
Slinn, W. G. N. 1974, Proceedings of the USAEC Symposium
Slinn, W. G. N. 1976, Geophys. Res. Lett., 3, 21
Squires, K. & Eaton, J. 1991, Phys. Fluids A, 3, 1169
Taylor, G. I. 1940, Collected Works G.I. Taylor, 3, 236
Turner, N. J., Fromang, S., Gammie, C., et al. 2014, Protostars and Planets VI,
References
Balachandar, S. & Eaton, J. K. 2010, Annual Review of Fluid Mechanics, 42,
Balbus, S. A. & Hawley, J. F. 1991, ApJ, 376, 214
Bec, J., Biferale, L., Boffetta, G., et al. 2006, Journal of Fluid Mechanics, 550,
111
349
411
Uhlmann, M. 2005, Journal of Computational Physics, 209, 448
Visser, R. G. & Ormel, C. W. 2015, ArXiv e-prints
Weidenschilling, S. J. 1977a, MNRAS, 180, 57
Weidenschilling, S. J. 1977b, Ap&SS, 51, 153
Weidenschilling, S. J. 1984, Icarus, 60, 553
Windmark, F., Birnstiel, T., Güttler, C., et al. 2012, A&A, 540, A73
Youdin, A. N. & Goodman, J. 2005, ApJ, 620, 459
H. Homann et al.: Effect of turbulence on collisions of dust particles with planetesimals in protoplanetary disks
Table A.1. Symbols used in this article
Symbol
Uc
I
Rep
Re
vL
u
Ek
εKol
νmol
d
Rp
(cid:96)Kol
tKol
L
tL
ts
S t
S tc
vc
α
νt
Σgas
Tgas
cs
r
z
H
Hd
ΩK
M(cid:63)
Description
average headwind velocity
turbulent intensity
planetesimal Reynolds number
outer gas flow Reynolds number
root-mean-square velocity of the gas flow
gas velocity
kinetic energy of the gas flow
mean kinetic energy dissipation rate
response or stopping time of dust
kinematic viscosity
planetesimal diameter
planetesimal radius
Kolmogorov length scale
Kolmogorov time scale
integral scale
integral time scale
Stokes number
critical Stokes number
dust collision speed
disk turbulence parameter
disk turbulent viscosity
disk surface density
gas temperature
speed of sound
orbital distance in the disk
vertical height in the disk
disk scale height
disk scale height for the dust
orbital (Keplerian) frequency
stellar mass
Equation
vL/Uc
Uc d/νmol
vLL/νmol
(cid:82)
√2/3Ek
(cid:82)
u2
∇ × u2
1
2
1
2
d/2
(ν3
mol/εKol)1/4
(νmol/εKol)1/2
v3
L/εKol
L/vL
ts/tc
νt/(cs H)
αcs H
(cid:17)−3/2
(cid:17)−1/2
(cid:16) r
(cid:16) r
au
au
kBTgas/µ
cs/ΩK
GM(cid:63)/r3
Σ1
T1
(cid:112)
(cid:112)
Article number, page 15 of 15
|
1003.1279 | 1 | 1003 | 2010-03-05T14:44:17 | Two new variable stars observed in the field of the extrasolar planet host star WASP-3 | [
"astro-ph.EP"
] | We report the discovery of two new short-period variable stars in the Lyra constellation, GSC2.3 N208000326 and GSC2.3 N20B000251, observed at the Astronomical Observatory of Autonomous Region of the Aosta Valley. Photometric measurements collected during several days are presented and discussed. One star appears to be a delta Scuti pulsating star (P=0.07848018+/-0.00000006 days; pulsation amplitudes V=0.055 mag and R=0.045; average (V-R)=0.378+/-0.009, probable spectral type F2). Th identity of the second star (P=0.402714+/-0.000008 days) resulted more difficult to be understood. We propose that this object should be classified as an eclipsing binary system where 0.065 and 0.055 are, respectively, the depths of the primary and secondary minimum in the light curve, as observed with a non standard R filter. | astro-ph.EP | astro-ph | January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
1, E. BERTOLINI
1,
TWO NEW VARIABLE STARS OBSERVED IN THE FIELD OF THE
EXTRASOLAR PLANET HOST STAR WASP-3
1, P. CALCIDESE
1, A. CARBOGNANI
2, A. BERNAGOZZI
1, P. GIACOBBE
M. DAMASSO
3, R. SMART
3, A. SOZZETTI
3
M. G. LATTANZI
1) Astronomical Observatory of the Autonomous Region of the Aosta Valley, Fraz. Lignan 39,
11020 Nus (AO) – Italy (www.oavda.it). E-mail of the corresponding author: [email protected]
2) Dept. of Physics - University of Turin, Italy
INAF - Astronomical Observatory of Turin, Pino Torinese, Italy
Abstract: We report the discovery of two new short-period variable stars in the Lyra constellation, GSC2.3
N208000326 and GSC2.3 N20B000251, observed at the Astronomical Observatory of Autonomous Region of
the Aosta Valley. Photometric measurements collected during several days are presented and discussed. One star
appears to be a δ Scuti pulsating star (P=0.07848018±0.00000006 days; pulsation amplitudes ∆V~0.055 mag and
∆R~0.045; (V-R)average=0.378±0.009, probable spectral type F2). The identity of the second star
(P=0.402714±0.000008 days) resulted more difficult to be understood. We propose that this object should be
classified as an eclipsing binary system where ~0.065 mag and ~0.055 are, respectively, the depths of the
primary and secondary minimum in the light curve, as measured with a non standard R filter.
3)
1
Introduction
Two new variable stars were observed at the Astronomical Observatory of the Autonomous
Region of the Aosta Valley (OAVdA; 45.7895° N, 7.47833° E). The OAVdA is located at
1675 m above the sea level in the Italian Alps, close to the border with France and
Switzerland. The discovery occurred during the ‘phase A’ of a photometric study aimed at
establishing the potential of the OAVdA site to host a long-term photometric transit search for
small-size (R<RNeptune) extrasolar planets (Damasso et al., 2009). As a part of this feasibility
study, during the period May-August 2009 and for a total of 19 nights we collected in- and
out-of-transit light curves of the star WASP-3 in the Lyra constellation, known to host an
extrasolar planet (Pollacco et al. 2007). These new variable stars represent an interesting by-
product of our observations of the WASP-3 field.
2
Instrumentation and methodology
All observations were carried out with a 250 mm f/3.8 Maksutov reflector telescope
mounting a CCD camera Moravian G2-3200ME with a camera sensor area of 2184 x 1472
square pixels (pixel linear dimension 6.8 µm) and a quantum efficiency of ~87% at
wavelength of 610 nm. This configuration results in a field of view of 52.10’ x 35.11’ and a
plate scale of 1.43 arcsec/pixel. All data were taken with an R filter (Astronomik) centered at
the wavelenght of 610 nm.
In order to collect standard photometric data for the two new variable stars, follow-up
observations were conducted using:
- the OAVdA largest optical telescope, an 810 mm f/7.9 Ritchey-Chrétien reflector
equipped with a Finger Lakes Instrumentation Pro Line PL 3041-BB back
illuminated CCD camera with standard BVRI filters. The camera sensor area is
2048 x 2048 square pixels (pixel linear dimension 15 µm), resulting in a field of view
of 16.3' x 16.3', with a plate scale of 1 arcsec/pixel in binning 2x2.
- a 400 mm f/7.64 Ritchey-Chrétien reflector telescope, mounting a front illuminated
CCD camera Finger Lakes Instrumentation Pro Line PL1001E with sensor
KAF1001E and standard BVRI filters. The camera sensor area is 1024 x 1024 square
1
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
pixels (pixel linear dimension 24 µm). The field of view is 26.4' x 26.4' with a plate scale of
1.55 arcsec/pixel.
To calibrate the V and R magnitudes we used the method described by Dymock and Miles,
2009 and magnitudes reported in the CMC14 (r’ magnitude) and 2MASS catalogues (J and K
magnitudes).
The data were reduced and elaborated with the automatic pipeline TEEPEE (Transiting
ExoplanEt PipElinE) that we have developed and tested as a part of our photometric study
focused on transiting extrasolar planets. TEEPEE performs image calibration (dark and flat-
fielding corrections, rotation and translation to the same pixel grid using a reference image),
airmass corrections (using an extinction coefficient averaged over typically dozens of stars)
and ensemble differential aperture photometry, using typically up to 100 reference objects
(depending on the target field) with apparent magnitudes close to the target magnitude.
To confirm that these variables are really a new discovery, we searched through several
sources: the General Catalogue of Variable Stars (GCVS,
http://www.sai.msu.su/groups/cluster/gcvs/gcvs/), the New Catalogue of Suspected Variable
Stars (http://www.sai.msu.su/groups/cluster/gcvs/gcvs/nsv/) and in VizieR database
(http://vizier.u-strasbg.fr/). Moreover, we also searched through the lists of new periodic
variable stars reported by the team of the SuperWASP project (Street et al., 2006; Norton et
al., 2008).
We adopted Sterken and Jaschek (1996), Breger (2000) and Percy (2007) as the main
references to suggest a classification for the new variable stars we observed.
3
Results and discussion on the individual objects
In Table 1 we summarize the positions, variability periods and amplitudes for both stars, and
in Table 2 we list the CMC14 stars we used for calibrating the V and R magnitudes. The
finding charts showing these stars are presented in Fig. 2 (1st variable) and Fig. 7 (2nd
variable).
3.1
Star #1 - GSC2.3 N208000326
In Fig. 1 we reproduce a finding chart for this object. This star is not reported as a variable
star in any of the catalogues we looked up. Figure 3 shows the phase-folded light curve for
this star (normalized flux) obtained using a pulsation period of 0.07848018±0.00000006 days,
estimated with the Lomb-Scargle algorithm implemented in the Starlink-PERIOD package
(Dhillon et al., 2001). The data presented here refer to 5 observing nights in the period second
half of July to second half of September 2009 and were collected with the 250 mm telescope
and R non standard filter. The error bars correspond to 1 σ. In Fig. 4 the R and V calibrated
light curves are showed, obtained with the 810 mm telescope during the follow-up
measurements of the night 28-29 September 2009. The plots show that the pulsation
amplitudes are ~0.055 mag in V and ~0.045 mag in R. Figure 5 shows the variation of the V-
R index recorded during the same night, with an average value (V-R)average=0.378±0.009. The
temporal variation appears to be in phase with V and R, strengthening the hypotesis that we
are in front of a pulsating star. To correct this value for the interstellar reddening, we used the
galactic dust reddening and extinction maps available on the Web at
http://irsa.ipac.caltech.edu/applications/DUST/ (the NASA/ IPAC Infrared Science Archive,
or IRSA) which are derived from data and techniques described in Schlegel, Finkbeiner &
Davis (1998). At the star location (galactic coordinates l=63.5919° and b=+18.4108°) the
2
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
V and R corrections are respectively AV=0.205 and AR=0.174 magnitudes (they represent
integrated value along the line of sight) corresponding to a V-R color excess E(V-R)=0.031.
This leads to the de-reddened estimate for the index color (V-R)average=0.347±0.009, which is
typical of a F2 type main sequence star, as reported in Zombek (1990).
Considering the shape of the light curve, the values found for the pulsation period, magnitude
variations and V-R index, we suggest that this star can be classified as a δ Scuti variable.
We can then estimate the mean absolute visual magnitude of the star using the Period-
Luminosity relation determined by Poretti et al. (2008), which does not depend on the
knowledge of the star metallicity:
MV = -3.65(±0.07)log (P) - 1.83(±0.08)
where P is given in days. We get MV = +2.20±0.11.
The mean apparent V magnitude (V=12.625±0.005 from our measurements) can be corrected
for the interstellar extinction using two different models of the galactic dust distribution,
which lead to two independent estimates for the star distance. Using the analytic expression in
Arenou et al. (1992), we assume for the interstellar extintion AV = 0.111 mag (integrated
along the line of sight) as estimated for a star with galactic coordinates l = 63.5919° and b =
+18.4108°. The maximum distance of the star results ~1153±26 pc (~3761 l.y.).
Using the IRSA galactic dust reddening and extinction maps we found AV = 0.205 mag for the
integrated galactic extinction along the line of sight at the star location. This leads to a second,
independent maximum star distance of ~1104±25 pc (~3600 l.y.).
Taking the average value of the two indpendent results, we then can assume d=1128.5±36 pc
as the better estimate for the maximum distance of the star.
3.2
Star #2 - GSC2.3 N20B000251
The finding chart containing the second variable star is showed in Figure 6. We did not find in
the literature any indication about the variability of this object. This target was observed on
the 250 mm telescope (and the Astronomik non standard R filter) for 9 nights in the period
June-July 2009. After the first inspection of the star light curve for the whole observing
period, the object showed what appeared a pulsation period of ~0.2 days (obtained applying
the Lomb-Scargle algorithm in the PERIOD package) and a magnitude amplitude of ~0.03
mag. These results in principle could indicate that the star is another δ Scuti. We then
measured V and R magnitudes on 14th October 2009 with the 400 mm telescope obtaining an
average V-R index (V-R)average=0.55±0.01, corresponding to a fraction of the estimated period
(~2 hr). The variations of the V-R index observed during that night is showed in Fig. 8. Then
we corrected this value for the effects of the galactic dust using the IRSA maps (galactic
coordinates of the target: l= 64.4724° and b=+18.5112°), and we assumed a color excess
E(V-R)=0.03±0.01. If the target were an isolated main sequence star, our best estimate of the
V-R index is indicative of a G-type star (Zombek, 1990). As δ Scuti are A-F spectral type
stars, we were forced to reconsider our hypothesis on the real nature of this variable.
We reorganized the data using P=0.402714±0.000008 days as the variability period, which is
double of the value we initially adopted. In Fig. 9 we show the corresponding normalized
phase-folded light curve. The error bars correspond to 1 σ. From the plot it results clear that,
using the new period, the object shows two distinct minima which differ only for ~0.01 mag,
while the maxima differs slightly in shape and havesimilar values. This clearly reveals that the
object is a binary system. To note that the light curve did not show evident variations in its
shape during the two-months period of our observations.
3
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
Only on the base of our photometric data some doubts necessairly remain regarding the
precise identity of the system, and we considered here two plausible alternatives.
The object could be a short-period non eclipsing binary system showing the properties typical
of an ellipsoidal variable (ELL variable star according the GCVS classification). In this case,
the small changes in the luminosty are due to the so called gravity darkening induced by
gravitational tidal distortion induced by the presence of a much less luminous compact object
as companion, which does not contribute to the observed radiative flux (e.g. Beech, 1985).
Besides, the system has a measured orbital period which falls in the range showed by
cataclysmic variables, close to the upper limit (e.g. Percy, 2007).
The second possibility is that the object is a short-period, grazing eclipsing binary system
where the primary and secondary stars could be assumed to be main sequence stars of
different spectral type.
Our opinion is that the latter hypothesis is much more reliable and it is based on the following
clues. In Table 3 we report the V, J, H and K magnitudes for this object as given by the
GSC2.3 and 2MASS catalogues. The V-K color index for this target is higher than expected
for a G-type star, while it is typical of a late K-type star (Zombek, 1990). If we are observing
an elongated star orbiting a compact object, the information we get from the V-R and V-K
color indexes, concerning the spectral type of the secondary star, is difficult to explain in
terms of just a difference in temperature between the hotter emisphere that points towards the
primary star (which is supposed to be colder) and the opposite. In this case, we should observe
a redder object (K-type) when the secondary star shows us the colder emisphere, and it should
appear as a G-type when it is in inferior conjunction. We did not find in literature examples of
secondary stars in close binary systems which show ellipsoidal variations associated with their
filled Roche lobe and which are characterized by an excursion in their spectral type as wide as
the one observed for our target.
We believe that the simplest and probably true scenario is assuming the system composed of
two normal stars, one that is cooler and brighter in red/near-infrared bands -then determining
the observed high value for the V-K index- while the hotter companion is brigther at shorter
wavelenghts, producing a V-R index shifted towards an upper spectral class.
As a further check to support our interpretation, we also looked up the HEASARC catalogue
(http://heasarc.gsfc.nasa.gov/, a very extended Web-based archive for gamma-ray, X-ray, and
extreme ultraviolet observations of cosmic sources) for a possible X-ray counterpart of the
target, giving hints of the presence in the binary system of a compact object and of an
exchange of matter with the secondary star, the donor companion. We did not find any source
at the variable star location which is associated to high-energy emissions. This evidence is in
favour of our proposed scenario.
Because we do not have any spectroscopic information on the object, we can not guess
anything more about the real nature of the stellar members of the system. Therefore we will
not go further in our discussion.
4
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
4
Conclusions
We reported the discovery of two new short-period variable stars in the Lyra constellation.
They represent interesting by-products of a campaing finalized to the observation of extrasolar
planets with the photometric transit method.
The objects were monitored for several days in order to better characterize them
photometrically and to guess the properties of their variability. According to our results, we
suggest that one star can be classified as a δ Scuti variable with asymmetric light curve (for
which we estimated the maximum heliocentric distance: d=1128.5±36 pc). The second object
should be classified as a detached eclipsing binary system but spectroscopic data are
necessary to definitively confirm this hypothesis.
5
Acknowledgments
This work has been possible thanks to the contributions of the European Union, the
Autonomous Region of the Aosta Valley and the Italian Department for Work, Health and
Pensions. We used the following resources: the VizieR catalogue access tool (CDS,
Strasbourg, France); the General Catalogue of Variable Stars (GCVS), mantained by the
Sternberg Astronomical Institute (Moscow, Russia); the Northern Sky Variability Survey,
operated by the University of California for the National Nuclear Security Administration
of the US Department of Energy; the NASA/ IPAC Infrared Science Archive, which is
operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract
with the National Aeronautics and Space Administration; the High Energy Astrophysics
Science Archive Research Center (HEASARC), provided by NASA's Goddard Space Flight
Center.
We thank Ennio Poretti (INAF-Osservatorio Astronomico di Brera) for the help in
understanding the nature of the second object discussed in this paper.
5
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
Table 1. Summary of the main data for the two new variable stars.
Star
number
Catalogue IDs
Coordinates
(J2000, GSC2.3
Catalogue)
Period
(days)
Magnitude
at max.
brightness
Magnitude
at min.
brightness
1
2
GSC2.3 N208000326
(also CMC14
183429.5+350424 ;
2MASS
18342951+3504242)
(α) 18:34:29.51
(δ) +35:04:24.3
0.07848018 ±
0.00000006
12.57±0.005
(Johnson-
Cousins V)
12.20±0.005
(Johnson-
Cousins R)
12.68±0.005
(Johnson-
Cousins V)
12.29±0.005
(Johnson-
Cousins R)
GSC2.3 N20B000251
(also CMC14
183528.6+355325 ;
2MASS
18352859+3553255)
(α) 18:35:28.62
(δ) +35:53:25.3
0.402714 ±
0.000008
13.69±0.01
(Johnson-
Cousins V)
13.15±0.01
(Johnson-
Cousins R)
Not
available
Suggested
variability type
(according the
GCVS)
DSCT
with asymmetric
light curve
E
Table 2. Data of the field stars used to calibrate the V and R magnitudes of the new variable stars, according
to the procedure described by Dymock and Miles, 2009.
Variable
star
number
Catalogue ID
for the field star
1
2
CMC14 183449.4+351240
CMC14 183428.4+350219
CMC14 183402.0+350401
CMC14 183451.4+350941
CMC14 183527.7+354601
CMC14 183550.1+354546
CMC14 183558.2+354717
CMC14 183532.9+354927
Coordinates
(J2000, CMC14
Catalogue)
(α) 18:34:49.477
(δ) +35:12:40.75
(α) 18:34:28.435
(δ) +35:02:19.81
(α) 18:34:02.082
(δ) +35:04:01.54
(α) 18:34:51.494
(δ) +35:09:41.13
(α) 18:35:27.794
(δ) +35:46:01.46
(α) 18:35:50.114
(δ) +35:45:46.56
(α) 18:35:58.238
(δ) +35:47:17.85
(α) 18:35:32.926
(δ) +35:49:27.39
r’ mag
(CMC14)
J-K index
(CMC14)
11.69
12.434
12.621
12.705
12.277
12.704
13.585
13.35
0.631
0.317
0.367
0.556
0.68
0.331
0.687
0.394
6
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
Table 3. Photometric data of the second variable star as given by the GSC2.3 (V magnitude) and 2MASS (J, H,
and K magnitudes) catalogues.
Va
J
H
K
V-Ka
13.57±0.36
11.870±0.021
11.358±0.029
11.274±0.025
2.296±0.360
a According our calibrated magnitudes, limited to ~2 hr of observations, we can assume V=13.73±0.01 as the better estimate for the average
V magnitude. This is in accordance with the value reported in the GSC2.3 catalogue.
If we assume our observed value for V and a color excess E(V-K)=0.158 as estimated using the IRSA galactic dust maps, the V-K color
index results to be 2.298±0.027, in agreement with the value reported in this Table.
Fig. 1. Image from the ESO Digitized Sky Survey database showing the star field of the
variable star GSC2.3 N208000326, which is indicated by a red arrow.
7
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
A
C
B
D
8
Fig. 2. Finding charts showing the field stars used to calibrate the V and R magnitudes of the first variable
star (indicated by a red arrow at the center of each image), as listed in Table 2. A: CMC14 183449.4+351240;
B: CMC14 183428.4+350219; C: CMC14 183402.0+350401; D: CMC14 183451.4+350941.
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
x
u
l
f
d
e
z
i
l
a
m
r
o
n
1.06
1.05
1.04
1.03
1.02
1.01
1
0.99
0.98
0.97
0.96
0.95
0.94
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
phase
Fig. 3. Normalized phase folded light curve of the variable star GSC2.3 N208000326, obtained using P=0.07848018 ±
0.00000006 days as the better estimate for the fundamental pulsation period (phase=0 corresponds to ephemeris
HJD=2455103.3939+P*E). The asymmetric shape of the light curve is clearly visible.
e
d
u
t
i
n
g
a
m
V
12.56
12.57
12.58
12.59
12.6
12.61
12.62
12.63
12.64
12.65
12.66
12.67
12.68
12.69
12.7
e
d
u
t
i
n
g
a
m
R
12.18
12.19
12.20
12.21
12.22
12.23
12.24
12.25
12.26
12.27
12.28
12.29
12.30
12.31
55103.33
55103.35
55103.37
55103.39
55103.41
55103.43
55103.33
55103.35
55103.37
55103.39
55103.41
55103.43
Modified HJD
Modified HJD
Fig. 4. V and R calibrated light curves of the pulsating star GSC2.3 N208000326 obtained during the night 28-29 September 2009, when
the object was monitored with the 810 mm telescope.
9
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
Fig. 5. Variation of the color index V-R observed for the star GSC2.3 N208000326 during
the night 28-29 September 2009.
Fig. 6. Image from the ESO Digitized Sky Survey database showing the star field
of the variable star GSC2.3 N20B000251, which is indicated by a red arrow.
10
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
A
C
B
D
11
Fig. 7. Finding charts showing the field stars used to calibrate the V and R magnitudes of the second variable
star (indicated by a red arrow at the center of each image), as listed in Table 2. A: CMC14 183527.7+354601;
B: CMC14 183550.1+354546; C: CMC14 183558.2+354717; D: CMC14 183532.9+354927.
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
0 .51
0 .52
0 .53
0 .54
R
-
V
0 .55
0 .56
0 .57
0 .58
0 .59
55119 .31
55119 .32
55119 .33
55119 .34
55119 .35
55119 .36
55119 .37
55119.38
55119 .39
55119 .40
Modified HJD
Fig. 8. Variation of the color index V-R observed for the star GSC2.3 N2B000251 during the night 14-15
October 2009.
x
u
l
f
d
e
z
i
l
a
m
r
o
n
1.06
1.05
1.04
1.03
1.02
1.01
1
0.99
0.98
0.97
0.96
0.95
0.94
-0.5
-0.4
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
phase
Fig. 9. Normalized phase-folded light curve of the variable star GSC2.3 N20B000251, obtained using P=
0.402714 ± 0.000008 days as the better estimate for the orbital period of the target (phase=0 corresponds to
ephemeris HJD=2454994.501335+P*E).
12
January 2010
OPEN EUROPEAN JOURNAL ON VARIABLE STARS
http://var.astro.cz/oejv
ISSN 1801-5964
References
• Arenou, F., Grenon, M. and Gómez, A., 1992, A Tridimensional Model of the Galactic Interstellar
Extinction, Astronomy and Astrophysics 258, 104-111 (1992A&A...258..104A)
• Beech, M., 1985, The Ellipsoidal Variables, Astrophysics and Space Science, Vol. 117, no. 1, pp. 69-81
(1985Ap&SS.117...69B)
• Breger, M., 2000, δ Scuti (review), in Delta Scuti and Related Stars, ASP Conference Series, Vol. 210, pp
3-42 (2000ASPC..210....3B)
• Damasso, M., et al., 2009, A Photometric Transit Search for Planets around Cool Stars from the Italian
Alps: Results from a Feasibility Study, to appear in Proceedings of the “Pathways Towards Habitable
Planets” meeting, Barcelona, September 14-18, 2009 (http://arxiv.org/abs/0911.3587)
• Dhillon, V.S., Privett, G.J., Duffey, K.P., 2001, PERIOD - A Time-Series Analysis Package, version 5.0
(http://www.starlink.rl.ac.uk/star/docs/sun167.htx/sun167.html)
• Dymock R., Miles R., 2009, A Method for Determining the V Magnitude of Asteroids from CCD Images,
(Journal of the British Astronomical Association, 119,3)
• Norton, A.J., et al., 2007, New Periodic Variable Stars Coincident with ROSAT Sources Discovered
Using SuperWASP, Astronomy and Astrophysics 467, 785-905 (2007A&A...467..785N)
• Percy, J.R., 2007, Understanding Variable Stars, Cambridge University Press
• Pollacco D., et al., 2007, WASP-3b: a Strongly Irradiated Transiting Gas-giant Planet, Monthly Notices of
the Royal Astronomical Society, Volume 385, Issue 3, pp. 1576-1584 (2008MNRAS.385.1576P)
• Poretti E., et al., 2008, Variable Stars in the Fornax dSph Galaxy. II. Pulsating Stars below the Horizontal
Branch, The Astrophysical Journal, Volume 685, Issue 2, pp. 947-957 (2008ApJ...685..947P)
• Schlegel, D.J., Finkbeiner D.P. & Davis, M., 1998, Maps of Dust IR Emission for Use in Estimation of
Reddening and CMBR Foregrounds, The Astrophysical Journal, Volume 500, pp. 525-553
(1998ApJ...500..525S)
• Sterken, C., Jaschek, C. (Ed.), 1996, Light Curves of Variable Stars. A Pictorial Atlas, Cambridge
Univerity Press
• Street, R.A., et al., 2007, SuperWASP-N Extrasolar Planet Candidates Between 18 < RA < 21h, Monthly
Notices of the Royal Astronomical Society, Volume 379, Issue 2, pp. 816-832 (2007MNRAS.379..816S)
• Zombeck M.V., 1990, Handbook of Astronomy and Astrophysics, 2nd edition, Cambridge University
Press
13
|
1804.10592 | 1 | 1804 | 2018-04-27T17:05:47 | On the orbital variability of Ganymede's atmosphere | [
"astro-ph.EP"
] | Ganymede's atmosphere is produced by radiative interactions with its surface, sourced by the Sun and the Jovian plasma. The sputtered and thermally desorbed molecules are tracked in our Exospheric Global Model (EGM), a 3-D parallelized collisional model. This program was developed to reconstruct the formation of the upper atmosphere/exosphere of planetary bodies interacting with solar photon flux and magnetospheric and/or the solar wind plasmas. Here, we describe the spatial distribution of the H$_2$O and O$_2$ components of Ganymede's atmosphere, and their temporal variability along Ganymede's rotation around Jupiter. In particular, we show that Ganymede's O$_2$ atmosphere is characterized by time scales of the order of Ganymede's rotational period with Jupiter's gravity being a significant driver of the spatial distribution of the heaviest exospheric components. Both the sourcing and the Jovian gravity are needed to explain some of the characteristics of the observed aurora emissions. As an example, the O$_2$ exosphere should peak at the equator with systematic maximum at the dusk equator terminator. The sputtering rate of the H$_2$O exosphere should be maximum on the leading hemisphere because of the shape of the open/close field lines boundary and displays some significant variability with longitude. | astro-ph.EP | astro-ph | On the orbital variability of Ganymede's atmosphere
F. Leblanc1, A.V. Oza1, L. Leclercq 2, 3, C. Schmidt1, T. Cassidy 4, R. Modolo.2, J.Y. Chaufray,2 and
R.E. Johnson3
1 LATMOS/IPSL, UPMC Univ. Paris 06 Sorbonne Universités, UVSQ, CNRS, 4 place Jussieu 75005
Paris, France
2 LATMOS/IPSL, UVSQ Université Paris-Saclay, UPMC Univ. Paris 06, CNRS, Guyancourt, France
3 University of Virginia, Charlottesville, Virginia, USA
4 Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, CO 80303, USA
1
Abstract : Ganymede's atmosphere is produced by radiative interactions with its surface, sourced by
the Sun and the Jovian plasma. The sputtered and thermally desorbed molecules are tracked in our
Exospheric Global Model (EGM), a 3-D parallelized collisional model. This program was developed
to reconstruct the formation of the upper atmosphere/exosphere of planetary bodies interacting with
solar photon flux and magnetospheric and/or the solar wind plasmas. Here, we describe the spatial
distribution of the H2O and O2 components of Ganymede's atmosphere, and their temporal variability
along Ganymede's rotation around Jupiter. In particular, we show that Ganymede's O2 atmosphere is
characterized by time scales of the order of Ganymede's rotational period with Jupiter's gravity being a
significant driver of the spatial distribution of the heaviest exospheric components. Both the sourcing
and the Jovian gravity are needed to explain some of the characteristics of the observed aurora
emissions. As an example, the O2 exosphere should peak at the equator with systematic maximum at
the dusk equator terminator. The sputtering rate of the H2O exosphere should be maximum on the
leading hemisphere because of the shape of the open/close field lines boundary and displays some
significant variability with longitude.
2
I Introduction
Ganymede is the only known satellite with an intrinsic magnetosphere (Kivelson et al. 1996). Its icy
surface is eroded by the solar radiation flux and by the Jovian energetic plasma leading to a locally-
collisional atmosphere and a locally surface-bounded exosphere (Stern 1999; Marconi 2007; Turc et al.
2014; Plainaki et al. 2015; Shematovich 2016). Unfortunately, Ganymede's atmosphere remains
essentially unknown despite the flybys during the mission Galileo in the 90s. This mission detected only
one atmospheric emission, Lyman α, interpreted as being produced by resonant scattering by exospheric
atomic hydrogen (Barth et al. 1997) with a column density around 9.21´1012 H/cm2. Such a column
density and its variation with distance to Ganymede suggested a surface density around 1.5´104 H/cm3
and an exospheric temperature around 450 K. Hubble Space Telescope was also able to identify another
component of Ganymede atmosphere, the O2 molecule through the observations of two UV atomic
oxygen emissions with intensities corresponding to a column density around 1 to 10´1014 O2/cm2 (Hall
et al. 1998). Since this discovery, several other HST observations of these emission have revealed their
auroral origin (Feldman et al. 2000; McGrath et al. 2013; Saur et al. 2015) highlighting the relations
between the spatial distribution of these emissions, Ganymede's intrinsic magnetospheric structure
(Kivelson et al. 1996) and Ganymede's surface's albedo (Khurana et al. 2007).
Gurnett et al. (1996) published the first observations of Ganymede's ionosphere by the plasma wave
spectrometer (PWS) on board Galileo suggesting ion densities on the order of 100 cm-3 and a scale
height around 1000 km. This first set of measurements was later complemented by a second flyby
(Eviatar et al. 2001b) confirming the density range but suggesting that a number density on the order of
1000 cm-3 at the surface may be more realistic and in better agreement with the upper limit set by radio
occultation (Kliore 1998). Measurements by the Plasma Science (PLS) instruments were also performed
during Galileo flybys in the polar regions of Ganymede (Frank et al. 1997). These authors reported the
observations of hydrogen ions outflowing from Ganymede's poles with temperatures around 4×104 K
and surface densities around 100 cm-3. These observations were later reanalyzed by Vasyliunas and
Eviatar (2000) suggesting that O ions rather than H ions constitute this outflow, with similar densities
as derived by Frank et al. (1997). These sets of observations of the thermal ionosphere were
3
complemented by energetic ion measurements obtained by EPD/Galileo, showing that Ganymede is
permanently embedded in an energetic plasma mainly composed of S+ and On+ (Paranicas et al. 1999).
After Galileo flybys, a simple 1-D model of the atmosphere and ionosphere was proposed by Eviatar et
al. (2001a) to fit PWS profiles with O2 densities at the poles around 106 cm-3. The first attempt to
reconstruct the origins and fate of Ganymede's neutral environment was described in Marconi (2007)
who developed a 2-D axi-symmetric model of this atmosphere using a combination of Direct Simulated
Monte Carlo and Test-particle Monte Carlo simulations focusing on the sunlit trailing hemisphere. He
suggested neutral densities peaking around 109 cm-3 around the subsolar region and 108 cm-3 at the poles
with the two main atmospheric species being H2O and O2. Turc et al. (2014) developed the first 3-D
model of Ganymede's atmosphere starting from Marconi (2007) simulation for the sunlit trailing
hemisphere. These authors found essentially the same surface density but some discrepancies were
found at high altitudes. Turc et al. (2014) found a Lyman α emission brightness up to 70 times smaller
than observed by Barth et al. (1997), whereas Marconi (2007) had to multiply his simulated H Lyman α
by a factor 4 to reproduce the observed intensities. Both models were able to reproduce the O2 emission
intensity observed by HST (Hall et al. 1998; Feldman et al. 2000). Furthermore, Plainaki et al. (2015)
developed a 3-D Monte Carlo model of Ganymede's exosphere, including a model of ion precipitation
in the auroral regions using a global MHD model of Ganymede's magnetosphere (Jia et al. 2009). They
modeled Ganymede's sunlit leading hemisphere highlighting the different exospheric regions of the two
main components H2O and O2.
In this paper, we improve Turc et al. (2014) model, implementing a collisional scheme in order to
identify the changes induced by atmospheric particle collisions. Marconi (2007) nicely illustrated the
remarkable nature of this exosphere/atmosphere by plotting his estimate of the Knudsen number around
Ganymede. Ganymede's atmosphere is predicted to be locally collisional (around the subsolar region)
and collisionless elsewhere. Another aspect of Ganymede that was not described in Marconi (2007) or
Turc et al. (2014), is how this atmosphere might change along Ganymede's orbit due to both its rotation
and Jupiter's gravitational field. These new simulations take into account the revolution of Ganymede
around Jupiter and Jupiter's gravitational field. In Section II, we present the main improvements of our
4
model with respect to Turc et al. (2014). Section III presents how this work could contribute to a better
understanding of Ganymede's atmospheric formation and evolution and is followed by a brief
comparison with the few observations of this atmosphere and with other similar observations in our
solar system in section VI.
II Exospheric Global Model
II.1 Application to Ganymede
A first version of the Exospheric Global Model (EGM) was applied to the description of Ganymede's
atmosphere in Turc et al. (2014). This is a 3-D Monte Carlo model describing the fate of test-particles,
ejected from a surface or an atmosphere of a planet or satellite, and subsequently followed under the
effect of gravitational fields until their escape (when reaching an upper limit set at few planetary radii
from the surface), their ionization/dissociation (by photo and/or electron impacts), or until their
readsorption at the surface (which is species dependent). In order to reconstruct density, velocity, and
temperature of the exospheric species (minor and major components), this model was parallelized
inducing a very significant reduction of the restitution time and the possibility to use much larger total
memory. These two improvements were crucial to describe the spatial and temporal distributions of all
exospheric species in synchrony, that is, in the case of Ganymede's icy surface, all H2O products, namely
H2O, H2, H, O2, O and OH. The list of the reactions taken into account in our model is given in Table 2
of Turc et al. (2014) and is reproduced in Table 1 below.
Reactions
H + hν ® H+ + e
H2 + hν ® H + H
+ + e
H2 + hν ® H2
H2 + hν ® H+ + H + e
O + hν ® O+ + e
OH + hν ® O + H
OH + hν ® OH+ + e
H2O + hν ® H + OH
H2O + hν ® H2 + O
H2O + hν ® H + H + O
Rate (s-1)
4.5×10-9
8.8×10-9
3.1×10-9
6.9×10-10
1.5×10-8
9.7×10-7
1.6×10-8
5.2×10-7
3.8×10-8
4.9×10-8
Excess Energy (eV)
References
3.8
2.2
6.9
26
24
3.4
22
3.4
3.4
4.6
Huebner et al. (1992)
Marconi (2007)
Huebner et al. (1992)
Huebner et al. (1992)
Huebner et al. (1992)
Marconi (2007)
Huebner et al. (1992)
Marconi (2007)
Marconi (2007)
Marconi (2007)
5
H2O + hν ® H + OH+ + e
H2O + hν ® H2 + O+ + e
H2O + hν ® OH + H+ + e
H2O + hν ® H2O+ + e
O2 + hν ® O + O
+ + e
O2 + hν ® O2
O2 + hν ® O + O+ + e
3.8×10-9
5.2×10-10
1.0×10-9
2.1×10-8
2.0×10-7
3.0×10-8
8.4×10-9
+ + e
Reactions
H + e ® H+ + e + e
H2 + e ® H + H + e
H2 + e ® H2
H2 + e ® H+ + H + e + e
O + e ® O+ + e + e
OH + e ® O + H
OH + e ® OH+ + e + e
H2O + e ® OH + H + e
H2O + e ® H2 + O + e
H2O + e ® OH + H+ + e + e
H2O + e ® H + OH+ + e + e
H2O + e ® H2 + O+ + e + e
H2O + e ® H2O+ + e + e
O2 + e ® O + O + e
+ + e + e
O2 + e ® O2
O2 + e ® O + O+ + e + e
9.1×10-8
9.6×10-9
1.6×10-8
9.6×10-10
2.0×10-8
1.2×10-8
2.8×10-8
3.7×10-8
1.6×10-8
4.3×10-9
4.0×10-9
7.1×10-9
2.1×10-8
1.3×10-8
2.0×10-8
1.1×10-8
References
Ip (1997)
Marconi (2007)
Ip (1997)
Ip (1997)
Ip (1997)
Ip (1997)
Ip (1997)
Ip (1997)
Ip (1997)
Ip (1997)
Marconi (2007)
Ip (1997)
Marconi (2007)
Ip (1997)
Smyth and Marconi (2006)
3.8
4.6
6.9
26
24
3.4
22
3.4
3.4
28
21
38
14
1.3
18
26
21
38
28
14
1.3
18
26
Huebner et al. (1992)
Huebner et al. (1992)
Huebner et al. (1992)
Huebner et al. (1992)
Marconi (2007)
Huebner et al. (1992)
Huebner et al. (1992)
Rate (cm3 s-1) Excess Energy (eV)
Table 1: Photon and electron ionization and dissociation in Ganymede's atmosphere (Turc et al. 2014)
Smyth and Marconi (2006)
When impacting the surface, test-particles can either stick or be re-ejected at once. As in Turc et al.
(2014) and Marconi (2007), electron ionization and dissociation impacts are described considering a
uniform mono-energetic electron in the Alfven wings of Ganymede's magnetosphere (open field line
regions, see section II.4) with an electron number density of 70 cm-3 and average energy of 20 eV. In
the closed field lines region, we neglected ionization and dissociation by electron impacts. We assume
that O2 and H2 species never stick to the surface, whereas all other species have a non-zero reabsorption
probability.
Jupiter's gravitational force at the surface of Ganymede is equivalent to a few percent of Ganymede's
gravitational force. However, because certain species have very long lifetimes in Ganymede's exosphere
(e.g. H2 and O2), Jupiter's gravity may play a significant role in shaping its exosphere, as will be shown
in Section III. Because we track molecules in Ganymede's reference frame, the Coriolis and centrifugal
6
forces act on Ganymede's molecules in its atmosphere throughout the orbit. This is only relevant for
certain species as the rotation period is ~7.2 days.
In Ganymede's inertial frame, the subsolar point rotates continuously, so that the day-to-night surface
temperature gradient drives a global migration of the exospheric species towards the nightside. On the
nightside, these particles can be readsorbed on the surface or thermally accommodate and be
immediately re-emitted generating a peak in atmospheric density near the surface, as observed on the
Moon (Hodges et al. 1974). We also introduce the possibility that returning volatiles can permeate the
regolith or be transiently adsorbed. These can eventually return to the exosphere, a process potentially
important for minor exospheric species, like Na, at Europa (Leblanc et al. 2005).
The standard outputs from this simulation are the 3-D values of the density, velocity, and temperature
of all species considered. We therefore define a 3-D grid with 100 cells in altitude distributed
exponentially from a cell of width 3 km in altitude at the surface to 300 km cell at 3 Ganymede radii
(RG), with 40 cells linearly spaced in longitude, and 20 cells in latitude, defined such that all cells have
the same volume at a given altitude. We also reconstructed 2-D column density maps as derived from
the 3-D neutral density, 2-D emission rates for the three main emission features observed around
Ganymede (Lyman α and 130.4 and 135.6 nm oxygen) and a 2-D map of the surface reservoir. The
model also calculates 3-D ionization rates by electron and photon impacts.
In the following, we describe the main improvements in the model of Turc et al. (2014), namely, a new
thermal surface temperature model of the surface (Section II.2), a new description of sublimation
(Section II.3), a new description of sputtering (Section II.4), the introduction of collisions (Section II.5)
and a fusion scheme for the test-particles (Section II.6).
II.2 Surface temperature
Ganymede's surface temperature is estimated using a one-dimensional heat conduction model in which
the thermal properties of the ice are treated as two layers (Spencer et al. 1989; available via
7
http://www.boulder.swri.edu/ spencer/thermprojrs). A low thermal inertia, 2.2×104 erg cm-2 s-1 K-0.5, is
assumed in the first few cm where the ice is heavily weathered, while the ice at depth has a much
larger thermal inertia of 7×104 erg cm-2 s-1 K-0.5, resulting in a longer thermal time scale. Density is set
to 0.15 and 0.92 g/cm3 in the top and bottom layers, respectively. The albedo is assumed to vary
linearly over the body's longitude, from 0.43 on the leading hemisphere to 0.37 on the trailing
(Spencer, 1987), and the emissivity is set at 0.96. A small correction for the latent heat of sublimation
is included (Abramov and Spencer, 2008). Solar insolation over a synodic day is determined in degree
increments of longitude with an eclipse length near equinox. This approach provides a basic
description of longitudinal asymmetries about the subsolar axis that result from the thermal inertia of
the ice, but does not account for any local variation of thermophysical parameters. Figure 1a shows the
subsolar temperature as a function of orbital longitude, which results from shadowed sunlight during
eclipse and albedo. Figure 1b maps the surface temperature at 270° orbital longitude (trailing) where a
temperature of 146° K is the diurnal maximum.
Figure 1: a: Evolution of the subsolar surface temperature with respect to Ganymede orbital longitude
(or phase angle). The insert in panel a is a zoom on the eclipse period showing the typical length of
the eclipse and its impact on the subsolar surface temperature. b: Map of the surface temperature at
270° Ganymede phase angle with respect to latitude and West longitude. The subsolar point is at a
latitude of 0° and a west longitude of 270° (white cross). Dawn and Dusk terminators are represented
by the vertical dashed lines and labeled accordingly.
8
As shown in Figure 1, the subsolar surface temperature peaks when Ganymede's trailing hemisphere is
illuminated and is minimum at leading because of the change in surface albedo. The surface
temperature displays a small but clear shift towards the afternoon due to lag from thermal inertia. The
afternoon side (between 270° to 0° west longitude) of the surface is therefore significantly hotter than
the morning side (between 180° and 270° west longitude).
II.3 Sublimation
In Marconi (2007), the sublimation rate at Ganymede's surface at a given temperature was given by
F = a × Ts
-0.5 × exp(-b/Ts)
[cm-2 s-1]
(2)
with a=1.1×1031 cm-2 s-1 K1/2 and b = 5737 K. Fitting the calculation by Johnson et al. (1981), we
found different values for a = 1.92×1032 cm-2 s-1 K1/2 and b = 6146 K. Fray and Schmitt (2009)
published another formulation for this sublimation rate, also used by Vorburger et al. (2015). We fit
the method of Vorburger et al. (2015) with equation (2) and derived values for a=2.17×1032 cm-2 s-1
K1/2 and b = 5950 K (section II.2). These three sublimation rates are shown in Figure 2 below. We also
considered a case where sublimation rate was significantly decreased (low sublimation case in Figure
2) to simulate an H2O exosphere dominated by sputtering. The Fray and Schmitt (2009) sublimation
model represents an ideal case, based on experiments for pure crystalline ice. In practice, sublimation
is highly dependent on the local grain structure and on an exact knowledge of the surface temperature,
an error of 10 K giving a few orders of magnitude difference in sublimation rate. We therefore
consider that even this low sublimation scenario might be within the realm of plausibility at
Ganymede.
9
Figure 2: Sublimation rate with respect to surface temperature as suggested by Fray and Schmitt (2009),
red solid line, by Johnson et al. (1981), cross black symbols and by Marconi (2007) in square green
symbols. The dashed blue line is the rate used for the low sublimation case throughout this paper.
In the results presented below, we will use the Low sublimation case (Sections III, III.1 and III.2) to
model the sublimation as well as Fray and Schmitt (2009) sublimation scenario (in Section III.3) to
illustrate how a strong sublimation rate could change the structure and content of the exosphere.
II.4 Sputtering
Sputtering is due to energetic Sn+ and On+ Jovian particles that penetrate Ganymede's intrinsic
magnetosphere through the polar regions and impact the polar surface regions (Cooper et al. 2001). This
bombardment leads to the ejection of surface species, in particular H2O, O, H, and OH, but also the
production of O2 and H2 by radiolysis in the regolith in which the molecules leave with near thermal
speeds. In order to describe the spatial distribution of this bombardment, in Turc et al. (2014), we
arbitrarily limited the sputtering to latitudes above 45° as did Marconi (2007). However, based on the
McGrath et al. (2013) auroral observations, we map out more precise regions where surface impacts
10
might preferentially occur, using latitudinal limits of the auroral regions corresponding to the footprints
of open field lines where Jovian plasma can penetrate through Ganymede's magnetosphere (Figure 3).
Figure 3: Position of the peak auroral emission brightness as reported by McGrath et al. (2013) in a
west longitude - latitude map (black symbols). The red and blue symbols represent the polynomial fit
performed in our model to reproduce the closed/open field line boundaries in each hemisphere. The
magnetosphere is most compressed in the ram direction at a longitude of 270°, corresponding to a
higher latitude boundary. The grey region corresponds closed Ganymede field lines where we assume
negligible sputtering.
By excluding the closed field line regions, we neglect any surface production of exospheric particles by
sputtering here, because according to Fatemi et al. (2016) and Leclercq et al. (2016), the total
precipitation rate in the closed field line region is much smaller than the total precipitation rate in the
opened field line region (see also Plainaki et al. 2015). However, ions reaching the closed field line
region are more energetic and should, on average, correspond to particles that sputter with higher yield,
an issue that has not been addressed in Fatemi et al. (2016) and Leclercq et al. (2016). In the following,
we neglect this contribution for simplicity. In the open field line regions, we assume a uniform incident
ion flux on the surface as suggested by Fatemi et al. (2016) and Leclercq et al. (2016). No variation of
11
the precipitating flux with respect to Jupiter plasma centrifugal equator or sheet has been included in
this modelling for simplicity.
To calculate the efficiency of an impacting ion to sputter surface particles, we include the surface
temperature dependence of the yield by Cassidy et al. (2013). Following their approach, we used
Johnson et al. (2009) and Fama et al. (2008) approximations of the yield (number of molecules and
atoms ejected per impacting particle) with respect to the incident energy:
YH2O(T) = YH2O(T0)× (1 + 220×exp(-0.06 eV/kB T)
(3)
and
YO2(T) = YO2(T0) × (1 + 1000×exp(-0.06 eV/kB T)
(4)
for sputtered H2O molecules (equation (3)) and O2 molecules (equation (4)). From these two equations,
and knowing the expected flux of ion impacting the surface, we calculate the sputtered flux from
Ganymede's icy surface. However, the value of YO2(T0) for O2 and YH2O(T0) for H2O needs to be
estimated as it depends on many parameters, such as the energy and charge of the impacting particle,
the surface structure, and the non-ice components (Teolis et al. 2016). Another approach is to choose
Y(T0) in order to reproduce the observed quantities in Ganymede's exosphere. For instance, the observed
FUV O2 auroral emission can be directly used to constrain Y(T0) (see also the discussion in Section IV).
Based on such observations, we can infer an effective yield of O2 molecules and then derive
corresponding effective yields for the other water products H, O, OH, H2, and H2O from their known
relative yields with respect to O2.
Marconi (2007) considered relative yield values equal to 1 for H2O, 1/18 for H, O, OH, to 16 for H2 and
to 8 for O2, whereas Smyth and Marconi (2006) suggested values equal to 1 for H2O, 1/17 for H, O, OH,
to 2 for H2 and to 1 for O2 at Europa. At Europa, Shematovich et al. (2005) considered that 1 molecule
of H2O should be ejected for each 1/10 O2 molecules. Noting the wide range in previous models, we
12
choose to base our yields on Cassidy et al. (2013), who made the first detailed calculation of sputtering
rates at Europa based on laboratory experiments, Europa's expected surface temperature, and the
expected intensity and spatial distribution of the impacting ions flux. These authors estimated that 2×1027
H2O/s and 1026 O2/s should be ejected from Europa's surface. In a companion paper, we applied the
EGM model to Europa's atmosphere (Oza et al. 2016) and thereby constrained Y(T0) for O2 and H2O
from the Cassidy et al. (2013) calculations of the total rates ejected from Europa's surface. According to
Oza et al. (2016), each time a molecule of H2O is ejected from the surface, an average of 1/20 O2 is
ejected from Europa's surface. The relative yield for H2 should be nearly stoichiometric, i.e. twice that
of O2, since they dominate the product composition and are both formed by radiolysis of the icy surface.
H, O, and OH yields should be small with respect to O2, H2 and H2O yield and are set to 1/20th of the O2
yield in agreement with Smyth and Marconi (2006). These yields represent a weighted average over all
charge states of O and S ions sourced in Io's plasma torus, which bears little compositional differences
between 9.4 and 15 Jovian radii (Delamere et al., 2005).
When an exospheric particle is ejected from the surface by sputtering, we proceed in the same way as
in Turc et al. (2014). Namely, H2O, O, H, and OH are ejected from the surface following an energy
distribution described in equation (8) in Turc et al. (2014) and following a Maxwell-Boltzmann
distribution at the local surface temperature in the case of O2 and H2 (equation 7 in Turc et al. 2014). In
order to properly describe the entire energy distribution up to the escape energy, we bin these two energy
distributions into 15-to-25 energy intervals (from 0 to an energy slightly above the escape energy) and
simultaneously eject 15 to 25 individual test-particles with a weight calculated as the integral of the
energy distribution over the energy range of each bin. Teolis et al. (2016) suggested that O2 molecules
could be ejected following the sum of two distributions: 70% of the O2 being ejected following a
Maxwellian Boltzmann distribution and 30% being ejected following a distribution in E/(U+E)3
(Johnson et al. 1983). Such a choice would populate the higher altitudes slightly more efficiently than
in our approach but would not change the main conclusions of this paper.
II.5 Collisional scheme
13
Because in some regions of Ganymede's atmosphere, collisions between atmospheric particles might
shape the density (Marconi 2007), we also include collisions between molecules and atoms. At each
time step, the total density of each species in a cell is calculated. From these densities, the maximum
number of collision pairs between each test-particle and the other test-particles for a given species in the
cell is calculated using the approach described in Bird (1994) with cross sections taken from Lewkow
and Kharchenko (2014). These authors developed a new scheme including cross sections depending on
both energy and angle that allows an accurate description of collision between any species colliding
with energies below a few tens of eV. Typical collision cross sections for molecule-molecule or
molecule-atom collision at 1 eV are 4.8×10-15 cm-2 and do not depend neither significantly on energy
between 0 to few eVs nor on the species considered in this work (H, H2, O, OH, O2 and H2O). These
cross sections are close to the ones used by Marconi (2007), as an example, equal to ~6.9×10-15 cm-2 for
H2O-H2O collision or to ~3.9×10-15 cm-2 for O2-O2 collision. We then randomly select a number of test-
particle pairs equal to the expected maximum number of collisions that should occur according to Bird
(1994). One of the difficulties when including minor species is that test-particles, in a given cell, might
have very different weights (i.e. they might represent vastly different number of real particles). We
therefore developed a new scheme to treat such collisions. For a test-particle A of weight WA and species
SA, we randomly determine with which species, SB, a test-particle A will most likely collide. We then
𝑊"#
$%#&'
randomly select nA test-particles Bi of weight WBi of the species SB in order to have
≥ WA. For
each pair of test-particles Bi and A, we then evaluate the likehood of a collision taking into account the
relative velocity between A and Bi. If a collision occurs, we determine new velocity vector after the
collision, V'Ai and V'Bi following Lewkov and Kharchenko (2014). We then create a new test-particle Ai
with velocity V'Ai and weight 𝑊(cid:0)# and repeat that scheme nA times. The weight of the particle A is then
reduced by WBi. The velocity of each Bi test-particle after collision is equal to V'Bi with the exception of
th Bi test-particle which is divided into two particles: one with velocity V'BnA and weight W'BnA =
and the other with velocity VBnA and weight WBnA - W'BnA.
𝑊"#
$%('
#&'
the nA
WA -
All collisions occur with energies below 10 eV so that molecular dissociation is negligible.
14
II.6 Fusion scheme
Since the collisional scheme implies the creation of new particles (see section II.5), we also had to
implement a fusion scheme in order to limit the number of test-particles followed by the model at a
given time. We tested two approaches to merge test-particles, both suggested by Lapenta (2002), the
binary scheme, and the tertiary scheme. The binary scheme is faster and since it leads to similar results,
we employ it. The binary scheme consists in selecting a pair of test-particles A and B which are close
each other in phase space (with positions qA and qB, velocities VA, VB and weight WA, WB respectively)
of the same species and in the same cell that minimizes (WA + WB)(VA² - VB²). We then create a new
particle with weight equal to W=WA+WB, position equal to q = (WAqA + WBqB)/(WA+WB) and velocity
equal to V = (WAVA + WBVB) / W.
For every arbitrarily fixed number of time steps (typically one hundred time steps), we count the number
of particles of each species within one cell. If this number is larger than a threshold, defined as to avoid
too many test-particles of a given species in a cell, the fusion scheme is activated, adapting to the
computational load this presents. Test-particles are therefore selected following the approach described
above, and are merged as long as this number continues to exceed the threshold.
III Ganymede's atmosphere
In the following, we will use the phase angle of Ganymede (equivalent to its sub-observer longitude)
to indicate its position around Jupiter. A phase angle of 0° corresponds to Ganymede being in the
shadow of Jupiter at midnight local time (eclipse), whereas phase angles of 90° and 270° correspond
to the sunlit leading hemisphere apex and sunlit trailing hemisphere apex respectively. Positions at the
surface are defined using west longitude (increasing with local time), 0° planetary longitude being the
subsolar point at a 0° phase angle. The simulation is performed in a GphiO rotating frame with the x-
axis always pointing along the corotational direction, y-axis pointing towards Jupiter, and the z-axis
pointing towards Ganymede's north pole (the x and y axis are inside the orbital plane of Ganymede).
Section III.1 is dedicated to the description of the fate of Ganymede's exosphere along its rotation
15
around Jupiter, without taking into account the effect of collision between atmospheric particles, an
issue specifically discussed in Section III.2. Section III.3 is based on simulations where collisions
were also neglected.
Using the sputtering description in section II.4, the typical surface temperatures of Figure 1, and the
low sublimation rate as described in Figure 2, the global sublimated flux is equal to 8.0×1021 H2O/s
whereas the average ejected flux by sputtering is equal to 8.0×1027 H2O/s, 3.6×1026 O2/s, 7.2×1026 H2/s
and 1.0×1025 H, O and OH/s. Marconi (2007) and Turc et al. (2014) used a flux of 1.5×1026 H2O/s and
1.2×1027 O2/s for sputtering, whereas Plainaki et al. (2015) used a flux of 7×1025 H2O/s for sputtering
and 7×1029 H2O/s for sublimation using the same model for sublimation as in Marconi (2007) and
Turc et al. (2014) (green square symbols in Figure 2).
We simulated more than 4.5 rotations of Ganymede around Jupiter in two weeks on 64 CPUs (that is,
3.7×106 time steps of 0.75 s each for a total of 780 simulated hours). During that run, 2.6×107 H2O
test-particles (simulating a total of 2.2×1034 H2O molecules) and 2.1×107 O2 test-particles (for 1034 O2
molecules) were ejected from the surface. Of these 1.2×106 H2O test-particles escaped the simulation
domain, 2.4×107 were reabsorbed in the surface, 6×105 were dissociated or ionized and none were
merged (meaning that the maximum number of test-particles in a cell allowed by the simulation was
never reached for H2O). For O2, 1.3×106 were dissociated or ionized, 1.9×107 were merged. The 3×105
O2 test-particles which escaped the simulation domain were generated from the tail of the energy
distribution ejected from the surface. Being associated with very small weight, these escaping
molecules represented an escaping flux of 6×1017 O2/s. Merging test-particles is much more important
for O2 than for H2O, because O2 test-particles are essentially confined near the surface. At each time, a
total number of 1 to 2×106 test-particles are followed. After convergence towards a steady state, the
typical escape rates are 1.6×1027 H2O/s, 6×1017 O2/s, 1.7×1026 O/s, 1.2×1026 OH/s, 3.0×1026 H2/s and
2.3×1026 H/s.
III.1 Seasonal variation of Ganymede's exosphere
16
A main improvement with respect to previously published models of Ganymede's atmosphere is that
we describe the evolution of Ganymede's exosphere throughout Ganymede's orbit around Jupiter while
taking into account the influence of Jupiter's gravitational field. By simulating a period of time long
enough to cover several Ganymede orbits, our model simultaneously calculates the motions of all
individual test-particles in Ganymede's inertial frame as well as the motion of Ganymede around
Jupiter. In general, only three orbits are needed to achieve a steady state exosphere on Ganymede.
Steady state is evaluated at a fixed phase between two successive orbits for a given time t, when the
reconstructed density, velocity, and temperature of the O2 and H2O exospheric components and their
loss rates vary by less than a few percent. In the results presented in this section, collisions are not
taken into account for the following results and will be specifically discussed in section III.2.
Most previous models have focused on one specific position of Ganymede corresponding to fixed
solar illumination and ram directions: the trailing illuminated hemisphere case in Turc et al. (2014)
and Marconi (2007), and the leading illuminated hemisphere case in Plainaki et al. (2015). However,
Ganymede's rotation implies that the geographical regions where sputtering and sublimation rates
reach a maximum, rotate with respect to one another. One direct consequence of this rotation is a
variable sputtering efficiency since yields depend strongly on the surface temperature and, therefore,
on the solar zenith angle (Cassidy et al. 2013).
The typical time for O2, ejected at the equator to migrate from one hemisphere to the other (that is to
move from a position at the surface to its opposite position on the other side of the satellite) can be
roughly estimated following Hunten et al. (1988). According to these authors, the collisionless
migration time of a particle moving in a surface-bounded exosphere is on the order of tm= tbl² where tb
= 2(H/g)1/2 is the ballistic hop time and l=RG/H. RG is Ganymede's radius and H is the scale height
H=kBTS/ (mg) with kB the Boltzmann constant, TS the surface temperature (as defined in section II.2),
m the mass of the particle and g the gravitational acceleration at the surface. For an O2 ejected from
noon at the sunlit trailing hemisphere to migrate across the hemisphere is, tm~765 hours or more than 4
orbital periods using Ts=143.5 K (see Figure 1a), H = 26 km, l=101, tb=270 s. Migration across the
hemisphere takes tm~1690 hours, or more than 9 orbits using 85K, H=15 km, l=171, tb=208 s. Precise
17
migration times are not very meaningful, since tm well exceeds the diurnal timescale of the thermal
wave, yet, the calculation is useful to conceptualize the atmosphere's diffusion. Given that the typical
lifetime of an O2 is on the order of one orbital period, newly ejected O2 roughly fixed in Ganymede's
reference frame moves therefore in local time due to Ganymede's rotation. The typical time scale for
the formation of the O2 exosphere can be estimated by dividing the average vertical column density
(2.4×1014 O2/cm2) by the average ejection rate (3.5×108 O2/cm2/s) or ~190 hours, that is, more than
one orbital period. Therefore, the O2 exosphere at a given orbital position will depend on its history
and evolution over about one Ganymede revolution. Because the H2O molecules are quickly
reabsorbed in the surface, the timescale of the H2O exosphere is much shorter than one rotational
period. In the case of H2, due to its low mass, the H2 exospheric will rapidly form across Ganymede
becoming spatially uniform on timescales much shorter than an orbital period and is, therefore, less
dependent on localized production.
18
Figure 4: Vertical column densities (log10 cm-2) of the H2O (panel a) and O2 exospheres (panel b) at
different Ganymede phase angles (90°, 210°, 270° and 0°) with respect to latitude (vertical axis) and
west longitude (horizontal axis) in the low-sublimation scenario. The subsolar point is indicated by the
black or white cross on each panel, the dusk and dawn terminators by vertical dashed black or white
lines. The ram and wake hemispheric centers are also indicated. Ganymede's direction of rotation is
indicated by the counter-clockwise arrow in the plane of Jovian co-rotation. Jupiter is at the center of
each figure.
In Figure 4, we display the vertical column densities of H2O (panel a) and O2 (panel b) at four
positions around Ganymede. Each figure corresponds to the average simulated density over a 6°
interval centered on the indicated phase angle value. This interval allows the simulation to reach a
reasonable signal-to-noise ratio for the reconstructed densities. The H2O vertical column density peaks
most of the time (except at 270° phase angle) on the plasma-wake hemisphere (also the leading
hemisphere) because H2O is predominantly produced by sputtering. In the low sublimation rate
scenario displayed here, sputtering should occur at lower latitudes on the wake hemisphere than on the
ram hemisphere (as suggested by the opened/closed field lines boundary in Figure 3), more H2O is
19
therefore ejected on the wake hemisphere than on the ram hemisphere. This magnetospheric effect,
when convolved with the strong thermal dependence of radiolysis (Equation 3) and sublimation
(Figure 2), effectively increases the H2O ejection rate at the sunlit leading hemisphere/plasma wake
(see also Plainaki et al. 2015). We therefore simulate a slightly denser H2O exosphere on the leading
illuminated hemisphere position (radial column density of 1.5×1013 H2O/cm2) than on the trailing
illuminated hemisphere position (radial column density of 1.3×1013 H2O/cm2). Observed from the
Earth, the average line-of-sight (LOS) column density at a phase angle of 90° over the hemisphere
disk would be equal to 3.4×1013 H2O/cm2, whereas it would decrease to 2.9×1013 H2O/cm2 at a phase
angle of 270° (see also Figure 6). H2O in the equatorial plane has densities smaller than 105 H2O/cm3
coming essentially from the polar region, whereas densities at the poles, are greater by two orders of
magnitude. Since H2O molecules efficiently stick to the surface, the polar H2O molecules do not
migrate efficiently towards the equatorial regions and remain confined around their ejection region.
The O2 exosphere is significantly different at the equatorial regions because the probability for O2 to
stick to the surface, even at coldest nightside surface temperatures, is low. Consequently, the O2
exosphere is spread across all longitudes and latitudes. There is however a clear dusk/dawn asymmetry
of the O2 spatial distribution with a peak in density at afternoon between j=90° to j=200°, Ganymede
local time and in the early night between j=90° to j=210°. The average radial column density varies
from 2.2×1014 O2/cm2 at a phase angle of 20° to a peak of 2.5×1014 O2/cm2 at a phase angle of 150°.
Both the asymmetry between dusk/dawn terminators and the variation of the exospheric content are a
clear result of the production and loss timescales of the O2 exosphere along Ganymede's orbit and can
be understood by the variation of the ejection pattern of the O2 particles as described below and in
Figure 5.
20
Figure 5: West longitude - latitude maps of the rate of O2 ejection (in O2/cm2/s) by sputtering at
various Ganymede phase angles. On each figure the X indicates the subsolar point whereas the
vertical dashed lines represent the terminators labeled accordingly as dawn and dusk. Also indicated
on each panel are the positions of the wake and ram hemispheres.
Due to the thermal inertia of the surface, the role of the ram direction for the sputtering and the shape
of the opened field lines region, there is a larger flux ejected close to the dusk than to the dawn when
averaging over an orbit. The peak O2 source rate occurs around a phase angle of 90° (at 5×108
O2/cm2/s) due to the variation in latitude of the opened/closed field lines boundary from the ram to
wake hemispheres. At a phase angle of 90° the largest open field line region (wake hemisphere) is also
where the surface temperature is highest resulting in an enhanced sputtering efficiency according to
equations (3) and (4). In contrast, at a phase angle of 270°, the opened field line region is smallest on
the dayside (the ram hemisphere), leading to a globally lower rate of ejection (3×108 O2/cm2/s).
Sputtering ejection reaches a minimum at a phase angle of 0° (2.4×108 O2/cm2/s) because of the
21
eclipse induced by Jupiter's shadow and the rapid decrease of Ganymede's surface temperature (Figure
1a). Since the modeled atmosphere is nearly stationary, a variation in the ejection or loss rate produces
a variation in the total content of the O2 exosphere. Therefore, the total O2 content reaches a minimum
around a phase angle of 20° due to the low ejection rate during the previous half of the rotation,
between phase angles 200° to 20° with a peak around a phase angle of 150° following the maximum in
ejection rate around a phase angle of 90°. That is, Ganymede's O2 exospheric total content decreases
during a portion of its orbit centered around a phase angle of 330° where the loss rate due to electron
impact ionization/dissociation is larger than the ejection rate. During this portion of the orbit,
Ganymede's decaying exosphere is a residual that formed earlier and conversely, during the following
portion of the orbit around a phase angle of 150°, the total content increases. The O2 exosphere also
appears slightly more extended on the dayside (Figure 4) due to larger surface temperatures. In our
model, O2 is mainly ejected from the higher latitudes by sputtering as explained in section II.4 (Figure
5). However, the vertical column density of the O2 clearly peaks at the equator (Figure 4b) and not at
high latitudes. The O2 diffusion time suggests that the molecules remain close to their source regions.
However, this does not take into account the effect of the centrifugal force which tends to drive
molecules equatorward. This force has an average intensity equal to few percent of Ganymede's
gravitational force at the surface and can therefore, shape the O2 exosphere as seen in Figure 6a. The
observed dusk/dawn asymmetry seen in Figure 4a leads also to larger column density on the right
(dusk) side of the exosphere as seen from the Sun in Figure 6a. This dusk/dawn ratio in the O2 column
density peaks at a value of 3.4 at a phase angle of 150°. Ganymede's exosphere near the leading
hemisphere portion of the orbit is formed by the production and spreading of the high latitude ejecta as
seen from Figure 5. The dusk/dawn asymmetries simulated here are also produced when simulating
the orbital evolution of Europa's O2 exosphere using the EGM (Oza et al. 2016), in accordance with
recent HST orbital observations of auroral features there (Roth et al. 2016)
The H exosphere is produced from the polar region by sputtering in our model. The column density, as
seen from the Sun shown in Figure 6b, reflects the source region and that the ejection energy primarily
leads to escape. Sputtered H2, on the other hand, is ejected at the local surface temperature at lower
22
energies, and with its higher mass a significant fraction remains gravitationally bound. Since it does
not stick, a nearly isotropic H2 exosphere results as shown in Figure 6c (despite electron and photon
impact dissociation of H2). In the low-sublimation scenario presented in Figure 6, H2O is primarily
ejected from polar regions with energies comparable to escape energy. Therefore H2O tends to
populate high altitudes with column densities peaking near the regions of sputtered production as
indicated in Figure 6d.
Figure 6: Reconstructed column densities (in log10 cm-2) of the O2 exosphere (panel a), of the H
exosphere (panel b), H2 (panel c) and H2O (panel d) at a phase angle of 270° as seen from the Sun.
We also tested if the formation of a surface reservoir by the readsorped H2O test-particles
subsequently released again could induce a change of the exosphere density. However, because H2O is
the main component of the surface, we did not notice any effect associated with the formation of an
exospheric reservoir of H2O in the surface.
III.2 The role of collisions
23
As suggested by Marconi (2007), Ganymede's atmosphere might be locally collisional. We therefore
investigated the effect of momentum transfer collisions on the global distribution of the exosphere.
Since the implementation of collisions, requires of the order of a few weeks on 128 CPUs, we only ran
simulations on one phase angle. Starting from the steady-state solution calculated in the collisionless
regime (as described in section III.1), we simulated the fate of Ganymede's exosphere, including
collisions (section II.5), during a time period long enough (few tens of hours) to reach an accurate
description of the exosphere up to ~1200 km in altitude. In Figure 7, we compare the density profile of
the two main components in Ganymede's exosphere with and without taking into account the effects of
collisions at a phase angle of 90°.
As shown in Figure 7, the main effect of collisions on the density profile is to enhance the population
of O2 at high altitudes. Indeed, most of the collisions occur either between O2 molecules or between O2
and H2O molecules, the two main species in the collisional region near the surface (within the first
tens kilometers). In this layer, collisions between other thermal O2 molecules have a negligible effect
on the O2 radial density distribution. However, O2 collisions with the more energetically sputtered H2O
possessing a high energy tail (~1/E²) can add significant energy to the O2 molecules, thereby
expanding the O2 exosphere.
24
Figure 7: Density profiles simulated by EGM without collision (blue lines) and with collision (red
lines) as calculated at a phase angle of 90°. Panels a and b are for H2O and O2 exospheric species
respectively at the subsolar point whereas panels c and d are for H2O and O2 exospheric species
respectively at the polar region. The density profile is calculated as the average density within 10° of
the subsolar point (panels a and b) and to the North pole (panels c and d).
We find that our results are in good agreement with the conclusions by Shematovich (2016) model of
Ganymede's exosphere. This author found that collisions caused the 10 to 100 km altitude region to be
populated by both O2 and H2O. Similarly, Marconi (2007) suggested that charge exchange between O2
+
ion and O2 molecules could enhance the population of O2 in the upper exosphere. Charge exchange was
also investigated by Dols et al. (2016) and by Luchetti et al. (2016) at Europa who both suggested that
charge exchange between O2
+ and O2 might be the dominant loss process for that satellite. Including
charge exchange accurately requires coupling between a magnetospheric and an exospheric model,
which is work in progress. The near surface density has a typical scale heights for O2 of 50 km which is
significantly larger than the expected scale height, 27 km, for an atmosphere in thermal equilibrium with
the surface at a maximum temperature of 146 K. In this model, O2 molecules are always ejected from
25
the surface following a Maxwell-Boltzmann distribution defined by the surface temperature. However,
as discussed in section III.1, O2 molecules are particularly sensitive to the non-inertial forces because of
their long residence time in the atmosphere. A large portion of the O2 molecules come from regions far
from the local surface leading to a globally uniform layer at the poles. This was illustrated in Turc et al.
(2014) where the O2 density at the poles decreased much faster (with scale height around 22 km) than
in the subsolar region.
III.3 The role of sublimation
For the Fray and Schmitt (2009) parametrization of sublimation, the total ejection rate would be equal
to 3.5×1029 H2O/s (comparable to the 8.0×1027 H2O/s sputtered and to the 7×1029 H2O/s sublimated in
Plainaki et al. (2015)). This implies that the H2O exosphere would be dominated by sublimation. This
is clearly illustrated by simulation results in Figure 8a which displays the exospheric density in the
equatorial plane, with the same format as in Figure 4a. The H2O density peaks as expected around the
subsolar region with a clear shift towards the afternoon due to the shift of the maximum temperature
on the dayside (Figure 1b). The eclipse period is particularly dramatic, since as shown by Turc et al.
(2014), the quick decrease of the surface temperature (Figure 1a insert) leads to an almost complete
disappearance of the H2O exosphere (characterized by a collapse of the density by more than 4 orders
of magnitude). The extended part of the H2O exosphere (in longitude and altitude with density larger
than 104 H2O/cm3) is due to the sputtering of polar regions by the incident Jovian plasma. It is clearly
more important on the wake side of the satellite because sputtering occurs at significantly lower
latitudes than on the ram side (Figure 3 and section III.1). As soon as a H2O molecule reaches the
nightside, it sticks efficiently so that the H2O exosphere is essentially confined on the dayside. In this
scenario, the exosphere at the equator is dominated by the H2O.
26
Figure 8: Panel a: Vertical column densities (log10 cm-2) of the H2O exosphere at different Ganymede
phase angles (90°, 210°, 270° and 0°) with respect to latitude (vertical axis) and west longitude
(horizontal axis) in the high sublimation scenario. The subsolar point is indicated by the white cross
on each panel, the dusk and dawn terminators by vertical dashed black or white lines. The ram and
wake hemispheric centers are also indicated. Ganymede's direction of rotation is indicated by the
counter-clockwise arrow in the plane of Jovian co-rotation. Jupiter is at the center of each figure.
Panel b: Reconstructed column densities (in log10 cm-2) of the H2O exosphere at a phase angle of 90°
(left panel) and 270° (right panel) as seen from the Sun.
27
For a large sublimation rate, the typical LOS H2O column density (as seen from the Sun) would peak
around the subsolar region with values between ~3.1 and ~8.5×1015 H2O/cm² (Figure 8b). The H2O
sputtered from polar regions is also apparent in the Figure 8b. The slight asymmetry of the column
density with respect to the subsolar point in Figure 8b is due to the maximum surface temperature
occurring in the afternoon. Since dissociation of H2O is a minor source of H, the column density of H
is very close (typically 4×1010 H/cm²) to that for the low sublimation rate case (in Figure 8b).
IV Discussion and Comparison with observations
Barth et al. (1997) reported a H column density of 9.21×1012 cm-2 in Ganymede's exosphere from the
measured Lyman α emission of 0.56±0.2×103 Rayleigh measured by Galileo during its G1 flyby (at a
phase angle of 180° and a solar zenith angle of 60° with a latitude of 21°). Such emission intensity was
later confirmed with HST (Feldman et al. 2000). The simulated H column density at that phase angle
is on the order of 6×1010 H/cm2 in the polar region and is extended down to latitudes of 20° because of
the particular shape of the open field line boundary at that particular phase angle. Our calculated
column density is therefore two orders of magnitude lower than the estimated column density by Barth
et al. (1997). A two order of magnitude error in the efficiency sputtering H atoms is difficult to
imagine based on measured sputtering yields (Johnson et al. 2009).
In combination with resonant scattering, H2O dissociated via electron impact could be an additional
channel producing Lyman α emission. Considering an electron with an energy near 20 eV, the typical
cross section for the production of Lyman α by electron impact on H2O is 0.43×10-18 cm2 (Makarov et
al. 2004). While trace ~keV hot electrons are a negligible contribution, an accurate knowledge of the
core electron energy is particularly important since this cross section increases by one order of
magnitude at 50 eV and peaks at 7×10-18 cm2 at 100 eV. Using the cross section at 20 eV, the typical
column density of 1014 H2O/cm2 in the low sublimation scenario (Figure 6) and 1016 H2O/cm2 in the
Fray and Schmitt (2009) sublimation scenario would lead to an emission of few Rayleigh and few
28
hundred Rayleigh respectively. The Fray and Schmitt (2009) sublimation scenario would therefore be
in rough agreement with Galileo-observed Lyman α emission. The low sublimation scenario could fit
the observation for 40 eV core electron temperatures if the sputtering yield was ten times larger. A
further potential source of Lyman alpha emission from Ganymede's exosphere is electron impact
dissociation of H2, also suggested by Marconi (2007). Ajello et al. (1995) published a cross section for
the production of Lyman alpha emission from electron impact dissociation of H2 peaking at 7×10-18
cm2 at 60 eV and equivalent to 5×10-18 cm2 at an electron energy of 20 eV. Using the typical 2×1014
H2/cm2 column density we simulated, this process would contribute less than a few tens of Rayleigh.
The typical column density of the O2 exospheric component has been reported by Hall et al. (1998) as
being between 2.4×1014 and 1.4×1015 O2/cm² when observing Ganymede's trailing hemisphere in June
1996 (phase angle around 270°) and by Feldman et al. (2000) as being equal to 0.3 to 5×1014 (October
30, 1998 with a phase angle between 290° and 300°). McGrath et al. (2013) also showed the spatial
distribution of the auroral emission at different orbital positions characterized by localized bright
regions with peak brightnesses between 100 to 400 Rayleigh in good agreement with the previous
reported observations. As stated in Section II.4, the O2 sputtering yield was tuned in order to reproduce
the reported column density. We therefore simulated column densities varying from 5 to 6.5×1014
O2/cm² averaged over the disk as observed from the Sun, and peaking to values of 2×1015 O2/cm² at
the equatorial limb (Figure 6a).
We also reconstructed the auroral emission associated with our simulated column density (Figure 9),
by considering electron excitation by a uniform mono-energetic electron population impacting the
sputtered O2 in the Alfven wings of Ganymede's magnetosphere (open field line regions displayed in
Figure 3) as the primary emission mechanism. We did not attempt to reproduce the ring shape
emission in Feldman et al. (2000), McGrath et al. (2013), and Saur et al. (2015). Furthermore, we did
not consider an accurate description of the electric currents in the open field line regions as in Payan et
29
al. (2015). The electron population used to calculate the emission brightness is the same as earlier: an
electron number density of 70 cm-3 and average energy of 20 eV. The emission rate coefficient for the
production of OI (5So) atoms from O2 emitting photons at 135.6 nm wavelength is set to be 1.1×10-9
cm-3 s-1 as in Hall et al. (1998). The 135.6 nm emission produced by direct excitation of the O atoms
by electron impact is also included, but is much smaller than the O2 source, in good agreement with
Hall et al. (1998). The goal of our calculation is to look for any potential signature in the observed
auroral emission induced by the dusk/dawn asymmetry of the O2 exosphere (Figure 4). As shown in
Figure 9, the simulated emission intensities vary between 10 and 200 Rayleigh in good agreement with
the observed values. More interestingly, at a phase angle of 90° (panel a), Ganymede's dusk exosphere
is simulated as being up to two times denser than at dawn consistent with the column density of the O2
exosphere as seen from the Sun being 2.3 times larger on moon's esatern/dusk hemisphere with respect
to the column density on the western/dawn hemisphere. This dusk/dawn asymmetry is visible in the
simulated emission brightness, the dusk limb emission at mid-latitudes being significantly larger than
the dawn limb emission. Remarkably, a similar asymmetry between the dawn and dusk limbs was
reported by HST (panel a of Figure 2 in McGrath et al. 2013). At a phase angle of 270° (Figure 9,
panel b), the simulated dawn/dusk asymmetry is much smaller (the west/dawn to east/dusk
hemispheric average column density ratio begin equal to 1.3), and does not induce a significant
emission asymmetry (Figure 9b), as is also seen with HST observations (panel c of Figure 2 in
McGrath et al. 2013). As indicated by HST observations near the trailing hemisphere (Fig. 9b) the
auroral emissions occur at higher latitudes than at leading (Fig. 9a) since we are facing the ram
hemisphere where the open/closed field lines boundary is also shifted to higher latitudes (Fig. 3). The
third position reported in McGrath et al. (2013) at a phase angle of 350° (their Figure 2b; near eclipse)
is characterized by two bright spots on the right limb. Our simulated image only marginally
reproduces such an intense asymmetry between right and left hemispheres (the right to left column
densities ratio being equal to 1.2). However, this observation was partially reproduced by Payan et al.
(2015) in their Figure 1 using a homogeneous exosphere coupled to a 3D MHD simulation of
Ganymede's magnetosphere, suggesting that these bright spots at the limb might be partly due to the
current system in Ganymede's magnetospheric Alfven wings. Payan et al. (2015) provided a similar
30
comparison for the two other observed positions (Figures 9a and b) and noted that the bright spot at a
phase angle of 105-108° (Figure 9, panel b) could not be reproduced. We suggest here, that this bright
spot may be a signature of the strong dusk/dawn asymmetry at the leading hemisphere/plasma wake.
Additionally, this dusk/dawn asymmetry could contribute to the almost systematic right to left
brightness asymmetry observed in the auroral emissions (McGrath et al. 2013; Saur et al. 2015).
Figure 9: Simulated images of the HST observed emission brightness reported in McGrath et al. (2013)
at three different phase angles as indicated on each panel with emission brightness calculated as
explained in the text.
Using the emission rate coefficient for the 130.4 nm emission, the typical ratio between 135.6 nm and
130.4 nm is equal to 1.8 with decreasing values at the edge of the emission region where the O/O2
relative abundance tends to increase. Hall et al. (1998) and Feldman et al. (2000) reported ratios between
1.2 and 3.2.
31
Teolis et al. (2016) suggested that the radiolysis efficiency to sputter O2 molecules from an icy surface
might be between 50-to-300 times less efficient than predicted by laboratory experiments (Teolis et al.
2010) or as calculated by Cassidy et al. (2013). In this paper, we scaled this efficiency to reproduce the
observed O2 column density and found that the ejected O2 rate should be around 3.4×1026 O2/s, assuming
a flux of Jovian ions precipitating on Ganymede's surface between 3.3 and 3.7×1024 particles/s for the
thermal component (Leclercq et al. 2016) and ~1024 particles/s for the high-energy component (Fatemi
et al. 2016), leading to an average sputtering yield of ~100. Teolis et al. (2016) estimated the yield for
ion sputtering on the icy satellites Rhea or Dione to be between 0.01 and 0.03, 4 orders of magnitude
less than the yields inferred in this paper. The temperature difference between Rhea, Dione, and
Ganymede could only account for a factor of 10 larger. Since an ion precipitation rate that is two orders
of magnitude larger is not realistic, we note that the Jovian plasma is more energetic with respect to the
Saturnian plasma (Teolis et al. 2016). Such a difference in energy could also lead to an additional order
of magnitude larger efficiency at Ganymede than at Dione or Rhea, leaving a difference of two orders
of magnitude with respect to the conclusions of Teolis et al. (2016). This difference is re-enforced by
noting that two orders of magnitude smaller O2 column density would place atomic products below the
detection limit of FUV oxygen aurorae observations by HST. Teolis et al. (2016) also suggested that
for reproducing Rhea's exosphere evolution, it was needed to introduce the notion of a shallow 1-cm
regolith allowing for the diffusion of O2 into the regolith and its shielding from the impacting ions. The
release of the O2 is then allowed when the regolith rotates into sunlight and induces the diffusion of the
O2 molecules back to the surface where they can be released by sputtering. Applied to Ganymede's case,
such a process would increase the reservoir of O2 available for ejection (by desorption or sputtering) and
therefore help reconcile Teolis et al. (2016) sputtering efficiency and the efficiency inferred by our
modelling. It would imply that more than 99% of the O2 originally produced by radiolysis accumulates
in the regolith over time and is released into the exosphere by thermal desorption and/or direct
sputtering. In such a scenario, the O2 exosphere would be much more extended in longitude and latitude.
We should also expect a much less asymmetric dawn to dusk terminator, in which the O2 exosphere
32
behaves like the H2O exosphere. Moreover, given that O2 is more efficiently adsorbed and trapped inside
the regolith than simulated here, a dawn enhancement would be favored by the release of trapped O2 at
sunrise and a decrease in reservoir content with increasing solar zenith angle. O2 trapping in the regolith
has actually been suggested by ground-based observations peaking at the trailing hemisphere (Spencer
et al. 1995), and confined at the equatorial regions (Calvin and Spencer 1997). It was shown by Johnson
and Jesser (1997) that O2 (and O3) bubbles can be trapped as voids in Ganymede's regolith with the
caveat that they can be destroyed should the incident ion flux be too high. The absence of O2 at the
poles, as well as the significantly less O2 at the leading hemisphere/plasma wake might be explained by
a more efficient eroding of the trapped O2 bubbles by the incident plasma.
V Conclusions
In this paper, we present the first model of Ganymede's exosphere taking into account Jupiter's
gravitational influence and the evolution of Ganymede's exosphere along its orbit around Jupiter. This
3D Monte Carlo model takes into account sublimation and sputtering as the main processes of
ejection, electron impact ionization and dissociation in the exosphere as well as a detailed interaction
between the exosphere and the surface. The role of Ganymede's magnetosphere is described in a
simple way which allows us to clearly distinguish signatures associated with the magnetosphere to
those associated with the formation and evolution of the exosphere. Ganymede's exosphere appears to
be highly stratified, displaying significant structure in terms of density and composition due to the
variation of the ejection mechanisms and molecular movement over its orbital period. In particular,
our results suggest that the shape of Ganymede's O2 exosphere at a given orbital position is highly
dependent on the production earlier in its orbit. We also showed that the H2O exosphere is very
different for the high sublimation rate and a low sublimation rate models. Moreover, even if
Ganymede's atmosphere is quasi-collisional, collisions are shown to only have a minor effect on the
exospheric density distribution.
33
In this work, we show that, by observing Ganymede's multi-species, water product exosphere along its
orbit, significant information on the way Ganymede interacts with its environment should be
accessible:
- The formation of the exosphere is a complex process driven by the impacting plasma and the
resulting space weathering of the surface, whose efficiency depends on the surface characteristics (e.g.
temperature, composition, structure) and on the Jovian plasma parameters (e.g. flux, energy and spatial
distribution at the surface). The spatial distribution for the lighter species of the sputtered exosphere
should directly reflect the spatial characteristics of the interaction between Ganymede and Jupiter's
magnetosphere. Moreover, the sputtered exospheric temporal variability and composition should be
highly dependent on the surface temperature and composition and could therefore be used to constrain
these two surface properties.
- The global 3-D spatial distribution of the exospheric density will depend on the influence of Jupiter
and Ganymede's gravitational fields. The fate of the ejecta is also species-dependent due to the species
dependent surface interactions. As an example, the dusk/dawn asymmetry in O2 highlighted in this
paper depends on the mobility across Ganymede's surface affected by absorption, accumulation and
re-emission from the regolith. The observed spectral emission signatures, such as the auroral
emissions, could be potentially used to characterize not only the Jovian-Ganymede interaction (Saur et
al. 2015), but also the spatial distribution of the exosphere.
- Ultimately, tracking the evolution of Ganymede's exospheric morphology and content along its orbit
should help constrain its atmospheric formation, and the interaction of its magnetospheric plasma with
the atmosphere and surface. The Galileo spacecraft showed the dichotomy in the surface albedo
between the trailing and leading hemispheres, which suggest that, Ganymede's exospheric source can
depend strongly on geographic longitude, an effect we did not consider in this work, but will
implement in the future. To better constrain the full structure of the exosphere we suggest full orbital
observations of Ganymede's FUV oxygen aurorae capable of indicating the evolution of Ganymede's
atmosphere.
34
Acknowledgement: This work is part of HELIOSARES Project supported by the ANR (ANR-09-
BLAN-0223) and ANR MARMITE-CNRS (ANR-13-BS05-0012-02). Authors are also indebted to the
"Système Solaire" program of the French Space Agency CNES for its support. Authors also
acknowledge the support of the IPSL data center CICLAD for providing us access to their computing
resources and data. Data may be obtained upon request from F. Leblanc
(email:[email protected]).
35
References
Abramov O. and J.R. Spencer, Numerical modeling of endogenic thermal anomalies on Europa,
Icarus, 195, 1, 378-385, 10.1016/j.icarus.2007.11.027, 2008
Ajello J.M., Kanik I., Ahmed S.M. and J.T. Clarke, Line profile of H Lyman α from dissociative
excitation of H2 with application to Jupiter, J. Geophys. Res. et al., 100, E12, 26411-26420, 1995
Barth, C.A., Hord, C.W., Stewart, A.I.F., Pryor, W.R., Simmons, K.E., McClintock, W.E., Ajello,
J.M., Naviaux, K.L. and Aiello, J.J.: Galileo ultraviolet spectrometer observations of atomic hydrogen
in the atmosphere of Ganymede, Geophys. Res. Lett., 24, 2147–2150, 1997
Bird, G. A., Molecular Gas Dynamics and the Direct Simulation of Gas Flows, Clarendon, Oxford,
England, 1994.
Calvin W.M. and J.R. Spencer, latitudinal distribution of O2 on Ganymede: Observations with the
Hubble Space Telescope, Icarus, 130, 505-516, 1997.
Cassidy, T.A., Paranicas, C.P., Shirley, J.H., Dalton, J.B., I.I.I., Teolis, B.D., Johnson, R.E., Kamp, L.,
Hendrix, A.R., Magnetospheric ion sputtering and water ice grain size at Europa, Planet. Space Sci.,
77, 64–73, http://dx.doi.org/10.1016/j.pss.2012.07.008, 2013.
Cooper, J.H., R.E. Johnson, B.H. Mauk and N. Gehrels, Energetic ion and electron irradiation of the
icy Galilean satellites, Icarus, 149, 133, 2001.
Dols, V., Bagenal, F., Cassidy, T., Crary, F., Delamare, P., Europa's atmospheric neutral escape:
importance of symmetrical O2 charge exchange, Icarus, 264, 387–397,
http://dx.doi.org/10.1016/j.icarus.2015.09.026, 2016.
Eviatar A., Vasyliunas V.M. and D.A. Gurnett, The ionosphere of Ganymede, Planet. Space. Sci., 49,
327-336, 2001a.
Eviatar, A., Strobel, D.F., Wolven, B.C., Feldman, P.D., McGrath, M.A., Williams, D.J., Excitation of
the Ganymede aurora. Astrophys. J. 555, 1013–1019, 2001b.
36
Fatemi, S., Poppe, A. R., Khurana, K. K., Holmstrm, M., Delory, G. T., On the formation of
Ganymede's surface brightness asymmetries: Kinetic simulations of Ganymede's magnetosphere:
plasma interaction with Ganymede, Geophys. Res. Let., 43, 4745,
http://doi.wiley.com/10.1002/2016GL068363, 2016.
Feldman, P.D., McGrath, M.A., Strobel, D.F., Warren Moos, H., Retherford, K.D., Wolven, B.C.,
HST/STIS ultraviolet imaging of polar aurora on Ganymede. Astrophys. J. 535, 1085–1090, 2000.
Fama, M., Shi,J., Baragiola, R.A., Sputtering of ice by low-energy ions, Surface Sci., 602, 156–161,
2008.
Frank, L.A., Paterson, W.R., Ackerson, K.L., Bolton, S.J., Outflow of hydrogen ions from Ganymede.
Geophys. Res. Lett., 24, 2151–2154, 1997.
Fray, N., Schmitt, B., Sublimation of ices of astrophysical interest: A bibliographic review. Planet.
Space Sci. 57, 2053–2080, http://dx.doi.org/10.1016/j.pss.2009.09.011, 2009.
Gurnett, D.A., Kurth, W.S., Roux, A., Bolton, S.J., Kennel, C.F., Evidence for a magnetosphere at
Ganymede from plasma-wave observations by the Galileo spacecraft. Nature 384 (6609), 535–537.
http://dx.doi.org/10.1038/384535a0, 1996.
Hall, D.T., Feldman, P.D., McGrath, M.A., Strobel, D.F., The far-ultraviolet oxygen airglow of
Europa and Ganymede, The Astrophysical Journal 499, 475, 1998.
Hodges, R. R. J., J. H. Hoffman, and F.S. Johnson, The lunar atmosphere, Icarus 21:415-426, 1974.
Hunten, D.M., Morgan, T.M., Shemansky, D.M., The Mercury atmosphere. In: Vilas, F., Chapman,
C., Matthews, M. (Eds.), Mercury. University of Arizona Press, Tucson, pp. 562–612, 1988.
Huebner, W.F., Keady, J.J., Lyon, S.P., 1992. Solar photo rates for planetary atmospheres and
atmospheric pollutants. Astrophys. Space Sci. 195, 1–289. http://dx.doi.org/10.1007/BF00644558.
Ip, W.-H., 1997. On the neutral cloud distribution in the saturnian magnetosphere. Icarus 126,
42–57. http://dx.doi.org/10.1006/icar.1996.5618.
37
Jia, X., Walker, R.J., Kivelson, M.G., Khurana, K.K., Linker, J.A., Properties of Ganymede's
magnetosphere inferred from improved three-dimensional MHD simulations. J. Geophys. Res. 114,
A09209. http://dx.doi.org/10.1029/2009JA014375., 2009.
Johnson, R.E., Lanzerotti, L.J., Brown, W.L., Armstrong, T.P., Erosion of Galilean satellites by
magnetospheric particles. Science 212, 1027–1030, 1981.
Johnson R.E., Boring J.W., Reimann C.T., Barton L.A., Sieveka E.M., Garrett J.W. and K.R. Farmer,
Plasma Ion-induced molecular ejection on the Galilean satellites: energies of ejected molecules,
Geophys. Res. Let., 10, 9, 892-895, 1983.
Johnson, R.E. and W.A. Jesser, O2/O3 microatmospheres in the surface of Ganymede, ApJ, 480, L79-
L82, 1997.
Johnson, R.E., Burger, M.H., Cassidy, T.A., Leblanc, F., Marconi, M., Smyth, W.H., Composition and
Detection of Europa's Sputter-Induced Atmosphere. In: Pappalardo, R.T., McKinnon, W.B., Khurana,
K.K. (Eds.), Europa. University of Arizona Press, Tucson, 2009.
Khurana, K. K., R. T. Pappalardo, N. Murphy, and T. Denk, The origin of Ganymede's polar caps,
Icarus, 191, 193–202, doi:10.1016/j.icarus.2007.04.022, 2007.
Kivelson, M.G. et al., Discovery of Ganymede's magnetic field by the Galileo spacecraft. Nature 384,
537–541, 1996.
Kliore, A.J., Satellite atmospheres and magnetospheres. Highlights Astron. 11, 1065-1069, 1998.
Lapenta, G., Particle Rezoning for Multidimensional Kinetic Particle-In-Cell Simulations, J. Comp.
Phys., 181, 317–337, http://linkinghub.elsevier.com/retrieve/pii/S0021999102971263, 2002.
Leblanc, F., Potter, A.E., Killen, R.M., Johnson, R.E., Origins of Europa Na cloud and Torus, Icarus
178,367–385, http://dx.doi.org/10.1016/j.icarus.2005.03.027, 2005.
38
Leclercq, L., Modolo, R., Leblanc, F., Hess, S., Mancini, M., 3D magnetospheric parallel hybrid
multi-grid method applied to planet plasma interactions, J. Comp. Phys., 309, 295,
http://linkinghub.elsevier.com/retrieve/pii/S0021999116000061, 2016.
Leclercq L., R. Modolo, F. Leblanc, P. Louarn and J.E. Walhund, Multi-species hybrid simulations of
Ganymede's magnetosphere Geophys. Res. Let., Submitted, 2016.
Lewkow N.R. and V. Kharchenko, Precipitation of energetic neutral atoms and induced non-thermal
escape fluxes from the Martian atmosphere, ApJ, 790-798, doi:10.1088/0004-637X/790/2/98, 2014.
Lucchetti, A., C. Plainaki, G. Cremonese, A. Milillo, T. Cassidy, Xianzhe J. and V. Shematovich, Loss
rates of Europa's tenuous atmosphere, Plan. Space Sci., 130, 14-23,
http://dx.doi.org/10.1016/j.pss.2016.01.009, 2016.
Makarov, O. P., J. M. Ajello, P. Vattipalle, I. Kanik, M. C. Festou, and A. Bhardwaj, Kinetic energy
distributions and line profile measurements of dissociation products of water upon electron impact, J.
Geophys. Res., 109, A09303, doi:10.1029/2002JA009353, 2004
Marconi, M.L., A kinetic model of Ganymede's atmosphere. Icarus 190, 155– 174, 2007.
McGrath, M.A., Xianzhe, J., Retherford, K., Feldman, P.D., Stroberl, D.F., Saur, J., Aurora on
Ganymede. J. Geophys. Res.: Space Phys. 118 (5), 2043–2054, 2013.
Oza A. V., F. Leblanc, C. Schmidt, R. E. Johnson, Origin and Evolution of Europa's Oxygen
Exosphere, Icarus, Submitted, 2016.
Paranicas, C., W. R. Paterson, A. F. Cheng, B. H. Mauk, R. W. McEntire, L. A. Frank, and D. J.
Williams, Energetic particle observations near Ganymede, J. Geophys. Res., 104, 17,459, 1999.
Payan, A. P., Paty, C. S., Retherford, K. D., Uncovering local magnetospheric processes governing the
morphology and variability of Ganymede's aurora using three-dimensional multiuid simulations of
Ganymede's magnetosphere, J. Geophys. Res., 120, 401, http://doi.wiley.com/10.1002/2014JA020301,
2015.
39
Plainaki, C., Milillo, A. , Massetti, S. , Mura, A., Jia, X., Orsini, S., Mangano, V., DeAngelis, E.,
Rispoli, R., The H2O and O2 exospheres of Ganymede: the result of a complex interaction between the
jovian magnetospheric ions and the icy moon, Icarus, 245, 306–319,
http://dx.doi.org/10.1016/j.icarus.2014.09.018, 2015.
Roth, L., J. Saur, K. D. Retherford, D. F. Strobel, P. D. Feldman, M. A. McGrath, J. R. Spencer, A.
Blöcker, and N. Ivchenko, Europa's far ultraviolet oxygen aurora from a comprehensive set of HST
observations, J. Geophys. Res. Space Physics, 121, doi:10.1002/2015JA022073, 2016.
Saur, J., Duling, S., Roth, L., Jia, X., Strobel, D. F., Feldman, P. D., Christensen, U. R., Retherford, K.
D., McGrath, M. A., Musacchio, F., Wennmacher, A., Neubauer, F. M., Simon, S., Hartkorn, O., The
search for a subsurface ocean in Ganymede with Hubble Space Telescope observations of its auroral
ovals. J. Geophys. Res., 120, 1715, http://doi.wiley.com/10.1002/2014JA020778, 2015.
Shematovich, V.I., Johnson, R.E., Cooper, J.F., and Wong, M.C., Surface-bounded atmosphere of
Europa, Icarus, 173, 480–498, 2005.
Shematovich, V.I., Neutral Atmosphere Near the Icy Surface of Jupiter's Moon Ganymede, Solar
System Research, 2016, 50, 4, 262–280, 2016.
Spencer J.R., Icy Galilean Satellite Reflectance Spectra: Less Ice on Ganymede and Callisto?,
ICARUS 70, 99-110, 1987.
Spencer, J.R., Lebofsky, L.A., Sykes, M.V., Systematic biases in radiometric diameter determinations.
Icarus 78, 337–354, 1989.
Spencer, J.R., Cavin, W.M., and M.J. Person, Charge-coupled device spectra of the Galilean satellites:
Molecular oxygen on Ganymede, J. Geophys. Res., 100, 19,049-19,056, 1995.
Smyth, W.H., Marconi, M.L., Europa's atmosphere, gas tori, and magnetospheric implications. Icarus
181, 510–526, 2006.
40
Stern, S.A., The lunar atmosphere: history, status, current problems and context, Reviews of
Geophysics, 37, 4, 453–491, 1999.
Teolis, B.D., Jones, G.H., Miles, P.F., Tokar, R.L., Magee, B.A., Waite, J.H., Roussos, E., Young,
D.T., Crary, F.J., Coates, A.J., Johnson, R.E., Tseng, W.-.L, Baragiola, R.A., Cassini finds an oxygen–
carbon dioxide atmosphere at Saturn's icy moon Rhea, Science, 330, 1813–1815, 2010.
Teolis B.D. and J.H. Waite, Dione and Rhea seasonal exospheres revealed by Cassini CAPS and
INMS, Icarus 272, 277–289, 2016.
Turc, L., Leclercq, L., Leblanc, F., Modolo, R., Chaufray, J.-Y., Modelling Ganymede's neutral
environment: A 3D test-particle simulation. Icarus 229, 157–169, 2014.
Vasyliunas, V.M., Eviatar, A., Outflow of ions from Ganymede: A reinterpretation, Geophys. Res.
Lett. 27, 1347–1349, 2000.
Vorburger A., Wurz P., Lammer H.n Barabash S. and O. Mousis, Monte-Carlo simulation of Callisto's
exosphere, Icarus 262, 14–29, 2015
41
|
0903.5137 | 1 | 0903 | 2009-03-30T07:25:22 | A new chronology for the Moon and Mercury | [
"astro-ph.EP"
] | In this paper we present a new method for dating the surface of the Moon, obtained by modeling the incoming flux of impactors and converting it into a size distribution of resulting craters. We compare the results from this model with the standard chronology for the Moon showing their similarities and discrepancies. In particular, we find indications of a non-constant impactor flux in the last 500 Myr and also discuss the implications of our findings for the Late Heavy Bombardment hypothesis. We also show the potential of our model for accurate dating of other inner Solar System bodies, by applying it to Mercury. | astro-ph.EP | astro-ph |
A new chronology for the Moon and Mercury
German Aerospace Center (DLR), Institute of Planetary Research, Rutherfordstr. 2, D-12489 Berlin
Dipartimento di Astronomia, Universit`a di Padova, Vicolo dell'Osservatorio 2, I-35122 Padova
Simone Marchi
[email protected]
Stefano Mottola
German Aerospace Center (DLR), Institute of Planetary Research, Rutherfordstr. 2, D-12489 Berlin
INAF, Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 3, I-35122 Padova
Gabriele Cremonese
Dipartimento di Geoscienze, Universit`a di Padova, via Giotto 1, I-35137, Padova
Matteo Massironi
and
Elena Martellato
INAF, Osservatorio Astronomico di Padova, Vicolo dell'Osservatorio 3, I-35122 Padova
ABSTRACT
In this paper we present a new method for dating the surface of the Moon, obtained by
modeling the incoming flux of impactors and converting it into a size distribution of resulting
craters. We compare the results from this model with the standard chronology for the Moon
showing their similarities and discrepancies. In particular, we find indications of a non-constant
impactor flux in the last 500 Myr and also discuss the implications of our findings for the Late
Heavy Bombardment hypothesis. We also show the potential of our model for accurate dating of
other inner Solar System bodies, by applying it to Mercury.
Subject headings:
solar system: general -- planets and satellites: Earth, Mercury, Moon
1.
Introduction
Craters are among the most spectacular surface
features of the solid bodies of the Solar System.
Cratering studies provide a fundamental tool for
the age determination of planetary and asteroidal
terrains. Since the beginning of the lunar explo-
ration, age estimates for the lunar terrains were
derived, followed by chronology models for the
other terrestrial planets. The development of the
lunar chronology greatly helped in interpreting
the evolution of the Solar System and in particu-
lar of our own planet, the Earth. Recently, thanks
to a fleet of new space missions (Mars Express
to Mars, MESSENGER to Mercury, and Kaguya
to the Moon, only to name a few), this field of
research entered a new exciting phase, where ac-
curate age estimates provide means for detailed
small-scales geological studies.
The method currently used for dating purposes
defines a chronology (crater surface density vs.
absolute age) for a reference body for which ra-
1
impactor sizes and velocities for any body in the
inner Solar System. Thus, the lunar calibration
can be exported with precision to any other body.
A drawback of the method, however, is that the
chronology of the reference body depends -through
the scaling law- on the physics of the cratering
process, which is as yet not fully understood. Fur-
thermore, the precision of the age estimate relies
on the accuracy of the dynamical model. Encour-
agingly, the adopted dynamical model is capable
of providing a good representation of the present
asteroid population and size distribution both in
the near-Earth space and in the main belt, with
a maximum deviation amounting to less than a
factor of 2 (Bottke et al. 2005b).
2. Modeling the cratering on the Earth-
Moon system
The present inner Solar System is continu-
ously reached by a flux of small bodies having
sizes smaller than few tens of km. A fraction
of these bodies (namely those having perihelion
< 1.3 AU) are called near-Earth objects (NEOs).
Contributors to this flux are represented by main
belt asteroids (MBAs) and Jupiter family comets
(JFCs). Both contributions have been modeled
by a number of authors and it is generally ac-
cepted that the MBAs constitute the main source
for the flux presently observed in the inner Solar
System (e.g. Morbidelli et al. 2002), therefore we
neglect the cometary contribution. This flux is
sustained by a few fast escape tracks, which are
continuously replenished with new bodies as a re-
sult of collisional processes and semimajor axis
mobility. The main gateways to the inner Solar
System are the ν6 secular resonance with Saturn
and the 3:1 mean motion resonance with Jupiter
(Morbidelli & Gladman 1998; Morbidelli et al.
2002). One of the most recent and accurate models
concerning NEOs formation has been developed
by Bottke et al. (2000, 2002), while Bottke et al.
(2005a,b) modeled the main belt dynamical evo-
lution. These models have been adopted in order
to estimate the properties of the impactor flux at
the Moon.
diometric ages are available for different terrains.
Then, on the basis of models predicting the im-
pactor flux ratio between the reference body and
another generic body, it is possible to estimate
the age of the latter. This method can be defined
experimental, since it develops a chronology of a
reference body for which two measurable quanti-
ties are available: absolute ages and number of
craters for selected areas. The Moon is the only
body of the Solar System that could fulfill both
requirements so far. This method has been de-
veloped and refined in the last 30 years by many
researchers and represents the reference for dat-
ing purposes (e.g. Hartmann et al. 1981; Neukum
1983). In this approach, the development of the
chronology for the reference body does not depend
explicitly on the physics of the cratering process.
However, such information becomes necessary -via
a cratering scaling law- in order to apply the Moon
chronology to other bodies or to infer the flux of
impactors from the observed cratering record.
A possible alternative -purely theoretical- ap-
proach could be based on the accurate estimation
of the time-dependent impactor flux for each tar-
get body, through the history of the Solar System.
With this method, a direct comparison between
the observed distribution of craters on a given ter-
rain and the isochrones produced by the model
would give its age. Although the knowledge of
the formation and evolution of the Solar System
greatly improved in recent years, such ambitious
goal is presently far beyond the capabilities of the
available models, due to the large uncertainties in
the early stages of its formation.
In this paper we introduce a third approach for
determining the chronology of objects in the inner
Solar System. This method, which can be con-
sidered as a hybrid of the first two approaches,
is based on the dynamical model by Bottke et al.
(2000, 2002, 2005a,b) which describes the forma-
tion and evolution of asteroids in the inner Solar
System.
In this framework, first the flux of im-
pactors on the Moon is derived, then the impactor
size distribution is converted into a cumulative
crater distribution via an appropriate scaling law,
and finally the radiometric ages of different regions
-the Apollo and Luna landing sites- are used for
the calibration of the lunar chronology. The main
advantage of our approach is that the adopted
model naturally describes the distributions of the
2
2.1. Modeling the impactor flux: the sta-
tionary case
The impactor size distribution for the Earth (he)
can be written as follows:
The flux of impactors may be described as a
differential distribution, φ(d, v), which represents
the number of incoming bodies per unit of im-
pactor size (d) and impact velocity (v). An im-
portant quantity -since it can be constrained from
observations- is the size distribution of the incom-
ing bodies, namely the number of impactors per
unit impactor size. Let h(d) be such distribution,
we then have: h(d) ≡ R φ(d, v) dv. Therefore,
without loss of generality, we can write:
φ(d, v) = h(d)f (d, v)
(1)
where f (d, v) is the distribution of the impact
velocities (i.e. the impact probability per unit im-
pact velocity) for any dimension d, and which ful-
fills the normalization constraint R f (d, v) dv ≡ 1.
We used the NEO orbital distribution model com-
puted by Bottke et al. (2000, 2002), generated by
integrating two sets of test particles placed into the
ν6 and 3:1 resonances (with 2×106 and 7×105 par-
ticles, respectively). For each particle they com-
puted the impact probability and the impact ve-
locity relative to the target body. Following the
work of Marchi et al. (2005), we computed the im-
pact velocity distribution (corrected for the gravi-
tational cross-section of the targets). We assumed
that the Moon revolves around the Sun with an
orbit identical to that of the Earth. Moreover,
we corrected the impact probability provided by
the Bottke et al model in order to account for
the different collisional life-time of the particles.
This is necessary because the orbital evolution of
the particles is coupled with their size through
√d (Farinella et al. 1998;
the relationship τ ∝
Bottke et al. 2005b), with τ being the particles'
collisional life-time.
In figure 1 we show the re-
sulting impactor velocity distribution f (d, v) for
the two extreme size bins used in our simulations
and for both the Moon and the Earth1.
1The overall shape of the distributions vaguely resembles a
Maxwellian distribution. It is interesting that the distribu-
tions have a large number of high velocity (i.e. > 20 km/s)
impactors. This distribution may help to explain some
anomalies found for some terrestrial impact craters.
In
particular, only 3 out of about 60 known craters with ages
< 60 Myr have been at present associated to tektite strewn
fields (French 1998). Since tektites are highly shocked
super-heated melts ejected during hypervelocity impacts
3
he(d) = Pehn(d)
(2)
where hn is the NEO differential size distri-
bution and Pe is the intrinsic collision probabil-
ity. Following Bottke et al. (2005b), we adopted
Pe = 2.8·10−9 yr−1. The distribution hn has been
derived from Bottke et al. (2002), and it can be
scaled to the Moon and other bodies. In the case
of the Moon, this is done considering the Earth-
Moon ratio of the impact probability as a function
of impactor size (Marchi et al. 2005). Let Σ(d) be
this ratio, then the Moon impactor size distribu-
tion becomes:
hm(d) = Pehn(d)Σ(d)
(3)
In figure 2 the cumulative impactor size dis-
tribution for both the Moon and the Earth are
shown. A similar approach can also be used for
evaluating the flux on other targets, like Mercury
(see section 6). In equation 1 the dependence of h
and f over time is neglected since we are consider-
ing a stationary problem, while eq. 2 and 3 strictly
describe the present flux. Arguments about possi-
ble variations in the flux intensity and shape have
been discussed since the late 70s, and this still re-
mains one of the most debated topics in the field
of lunar cratering.
In their work, Neukum and
coworkers (e.g. Neukum 1983; Neukum & Ivanov
1994) generally assume a constant shape of the
impactor size distribution over time, while others
claim a sudden change in the shape after the Late
Heavy Bombardment (LHB), about 3.8 Gyr ago
(e.g. Strom et al. 2005). One of the potentials of
our model is that we can implement and evaluate
the effects of a time-dependent size-distribution
on the crater cumulative distribution, as shown in
section 5.
2.2. The scaling law
In order to convert the flux of impactors into
a resulting crater distribution, we need to apply
(e.g. Glass 1990; Koeberl et al. 1996), from simple statis-
tical arguments we may estimate that a threshold perpen-
dicular impact velocity of at least 24 km/s is required for
their formation (more than 33 km/s modulus for a π/4 im-
pacting angle).
a so-called crater-scaling law. Such scaling law
attempts to describe the outcome of a cratering
event based on the impact energy and on the phys-
ical properties of both the target and the projec-
tile. Once an appropriate scaling law is chosen, we
may write the final crater diameter D as a function
of the impactor diameter, impact velocity, and a
number of parameters (density, strength...
indi-
cated by ~p), namely D = S(d, v, ~p). Despite the
great effort (both computational and experimen-
tal) devoted to this task, the physics of crater for-
mation is far from being completely understood.
For this reason, we decided to explore how the pre-
dicted crater distribution depends on the choice of
the scaling law. For this purpose we considered the
three scaling laws most used in this field. The first
one by Holsapple & Housen (2007) (H&H here-
inafter), reads:
D = kd(cid:20) gd
2v2
δ(cid:19)
⊥(cid:18) ρ
2ν
µ
+(cid:18) Y
ρv2
⊥(cid:19)
ν(2+µ)
µ
2+µ
2 (cid:18) ρ
δ(cid:19)
2+µ
(cid:21)− µ
(4)
where g is the target gravitational acceleration,
v⊥ is the perpendicular component of the impactor
velocity, δ is the projectile density, ρ and Y are
the density and tensile strength of the target. The
quantities k and µ depend on the target material
and ν on its porosity. H&H estimated these latter
parameters by best-fitting over a range of exper-
iments done with different materials. From this
study we adopted the values k = 1.03 and µ = 0.41
for cohesive soils, and k = 0.93, µ = 0.55 for rocks.
We used ν = 0.4 in all cases (H&H). Equation 4
accounts for the transition between strength and
gravity regime, allowing a smooth transition be-
tween those extreme conditions.
The second scaling law considered is reported by
Ivanov et al. (2001), (adapted from Schmidt & Housen
1987) (I&S hereinafter):
D
d(δ/ρ)0.26v0.55
⊥
=
1.28
[(Dsg + D)g]0.28
(5)
where Dsg represents the strength-gravity tran-
sition crater. Dsg has been set equal to 120 m and
30 m for the Moon and the Earth, respectively
(Asphaug et al. 1996).
Finally, we have considered the scaling law
used by Stuart & Binzel (2004) (adapted from
Shoemaker et al. 1990) (S&S hereinafter):
4
D = 0.0224(cid:18)W
δ
ρ(cid:19)0.294(cid:18) ge
g (cid:19)1/6
(sin α)2/3
(6)
where, W is the impactor kinetic energy, ge is
the Earth gravity and α is the impactor angle with
respect to the surface (vertical impacts correspond
to α = π/2). The numerical multiplicative factor
takes into account the correction from transient
to final crater dimension (Stuart & Binzel 2004).
The crater size D reported in these formulas
should be regarded as the size of the final crater.
However, following Pike (1980), a correction for
the transition between simple and complex craters
has also been applied to all the scaling laws in the
form of:
D = D
D =
D1.18
D0.18
⋆
if D < D⋆
if D > D⋆
(7)
(8)
where D⋆ is the diameter of the transition from
simple to complex crater2. All impacts are as-
sumed to occur at the most probable impact an-
gle, namely π/4 with respect to the normal to the
surface.
All the scaling laws assume that the density and
the strength of the target are constant through-
out the body. However, most target bodies are
characterized by a layered structure, with a den-
sity increasing with depth. Therefore, in order to
produce a better representation of the target den-
sity, we have computed the average density over a
depth of about 10 times the radius of the projec-
tile. This is because the size of the crater is -on
average over different impact conditions- about a
factor ten larger than the size of the impactor.
In particular, in the case of the Moon we have as-
sumed a 10-km layer of fractured silicates (megare-
golith and heavily fractured anorthosites), on top
of a bulk anorthositic crust in turn laying above a
peridotitic mantle (see fig. 3). For the Earth (see
fig. 3) we considered the following lithospheric
structure (Anderson 2007): a 2 km thick layer
of sedimentary rocks, a mainly granitoid upper-
crust down to a depth of 20 km, a denser lower
2D⋆ has been set equal to 18 km and 4 km for the Moon
and Earth, respectively (Ivanov et al. 2001).
crust down to 40 km, a peridotitic lithospheric
mantle which is in turn characterized by an upper
layer with an average density of 3.2 g/cm3 (up to
150 km), and a lower layer with an higher density
(3.3 g/cm3). The impactor density has been set to
2.7 g/cm3. Regarding the strength, the value re-
ported by Asphaug et al. (1996) for bulk silicates,
namely Y0 = 2 × 108 dyne/cm2 was used both for
the Moon and Earth crusts. However, we assume
a linear increase of the strength, from zero at the
surface up to Y0 at the bottom of the heavily frac-
tured layer for the Moon in order to take into ac-
count the cohesionless regolith layer on the surface
and the underlying megaregolith and anorthosites
which are likely characterized by a progressive de-
crease in fracture density with depth. A similar
assumption was adopted also for the layer of sed-
imentary rocks on the Earth. The strength at the
Earth's surface was set to 0.3 ×108 dyne/cm2.
Following our description of the Moon crust, its
upper layers are mainly made of highly fractured
materials, which behave like cohesive soils. There-
fore, we set a sharp transition in the scaling law,
from cohesive soil in the case of small impactors
to hard rock for larger ones, at a projectile size
of 1/20th of the thickness of the heavily fractured
silicate layer (i.e. 0.5 km). For the Earth only the
hard rock scaling law was used since only rocky
layers have been assumed at the surface.
3. The Model Production Function (MPF)
We have now all the necessary inputs for com-
puting the cumulative crater size distribution
(hereafter the Model Production Function, MPF)
for the target bodies. Let Φ(D) be the crater
differential distribution, we may write:
MPF(D) = Z ∞
D
Φ( D)d D
(9)
where Φ can be expressed in terms of φ and S
(see appendix for details).
In figure 4 our lunar
model production function is shown for different
scaling laws along with the Neukum production
function for the Moon (NPF). The NPF has been
the traditional reference for dating purposes since
the late 70s, with a few revisions in more recent
times (Neukum 1983; Neukum et al. 2001). An al-
ternative production function, the so-called HPF,
has also been proposed by Hartmann et al. (1981).
5
A detailed comparison between these two pro-
duction functions can be found in Neukum et al.
(2001). For our purposes, we recall that the NPF
shows an overall agreement with the HPF, even
if some discrepancies are present. Nevertheless,
as stated in Neukum et al. (2001), the NPF-based
chronology represents the current best interpreta-
tion for the lunar cratering chronology. Moreover,
since the NPF extends to a wider size range than
the HPF, the former is more suitable for a thor-
ough comparison with the MPF. For these reasons,
here we provide detailed comparison between the
MPF and the NPF. In the following paragraph,
we briefly recall how the NPF has been derived.
There are several versions of the NPF. In this pa-
per we refer to the old NPF (Neukum & Ivanov
1994) and to the new NPF (Neukum et al. 2001).
In both cases, the underlying assumption is that
the shape of the production function has been
constant, with only the absolute number of craters
changing over time. Therefore, the NPF was built
by vertically re-scaling crater counts from areas
with different ages, until they overlapped. The
final curve was then expressed as a polynomial
fit of the re-scaled data points. The strength
of such procedure is that it is not model depen-
dent because based on measured data. On the
other hand this procedure, relying on the accu-
racy of the measurements in multiple overlapping
regions, is particularly prone to severe error prop-
agation, especially at the large diameter end. In
addition, in some regions the cumulative distri-
butions of craters can be altered by surface pro-
cesses like sporadic magma effusions, ejecta from
other craters etc. The subsequent effects of su-
perposition of geological units of different ages,
normally not recognizable on remote sensing im-
ages, can lead to mixed crater size distributions
(Neukum & Horn 1976). Figure 4 clearly shows
how the change of crater counting on single regions
may alter the final shape of the NPF. In partic-
ular, the discrepancies between the old and the
new NPF are due to an adjustment of the crater
counts for Orientale Basin (Ivanov et al. 2001).
A few considerations can be drawn from fig. 4.
The first one is that different scaling laws produce
differences on the MPF as large as a factor of ten
for large craters. We find that the MPF produced
with the S&S scaling law is not in good agreement
with the NPF throughout the whole size range.
When applying the same scaling law to the ob-
served NEO population, Stuart & Binzel (2004)
found a better match to the new NPF for craters
> 10 km. This discrepancy is mainly due to the
different value of the intrinsic probability that
Stuart and Binzel used in their model (namely,
1.50 · 10−9 yr−1). However, we believe the Bottke
et al. determination of the intrinsic collision prob-
ability to be more accurate since it relies on a bet-
ter treatment of the gravitational focusing effects
(A. Morbidelli, pers. communication). The main
difference between the H&H and the I&S scaling
laws takes place at crater sizes smaller than 10 km.
This can be traced back to our choice of using the
cohesive soil scaling law for small crater sizes in
the H&H scaling law which causes an inflection
point in the MPF. This transition in the target's
crustal properties produces an S-shaped feature
in the MPF, which is qualitatively similar to the
one found in the NPF. This finding would suggest
that the observed S-shaped feature in the NPF is
caused by physical properties of the target, rather
than reflecting the shape of the size distribution
of the impactors, as proposed by Werner et al.
(2002) and Ivanov et al. (2002). Actually, a weak
wavy feature in the relevant size range is present
in the hn(d) which, however, is smeared out when
passing to the MPF. However, in order to finally
assess this issue, a better estimate of the observed
impactor size distribution for sizes < 0.1 km is
needed. It should also be mentioned that the de-
tailed shape of the NPF somewhat depends on the
fitting procedure used to construct the NPF itself,
as shown by the mismatch between the old and
new NPF for large craters.
In conclusion, despite of all the caveats mentioned,
it is a remarkable result that the present MPF ob-
tained with the H&H scaling law and the NPF
(both new and old) are similar within a factor of
two throughout the whole size range considered,
that is from 0.01 km to a few hundred km. On
the other hand, the S&S and I&S scaling laws fail
to accurately reproduce the absolute number of
craters per km2 per yr, with a maximum devia-
tion of a factor of 10. The good agreement of the
absolute density of craters per yr derived from the
observations (either the NPF or the HPF) and
the MPF with the H&H scaling law is therefore a
clear indication that the latter scaling law is the
more appropriate in describing the lunar crater-
ing. This agreement is particularly interesting,
considering that the NPF and MPF are derived
from completely independent methods: one being
based on crater counting, and the other being the
result of theoretical modeling. For these reasons,
we decided to restrict the following analysis to the
H&H scaling law. We refer to this MPF as to the
nominal model.
3.1. The reliability of the nominal MPF
In this section we explore the effects that the
various parameters involved in our model have
on the nominal MPF described in the last sec-
tion.
In particular, we analyze the scaling law,
the NEO size distribution and the assumed den-
sity and strength profiles.
The H&H scaling law depends on three parame-
ters (namely k, ν and µ) which have been derived
by best fit to laboratory data. We applied a varia-
tion of 10% around their nominal values, and com-
puted the resulting MPF. It results that k and ν
have a negligible effect on the MPF. Changing µ of
±10%, the MPF shifts vertically of about a factor
of ±2, respectively.
A detailed comparison of the modeled and ob-
served NEO size distributions can be found in
Bottke et al. (2005a,b). For our purposes, we no-
tice that they basically overlap for d > 1.5 km.
For smaller impactor sizes, the situation is less
clear since different surveys produce slightly differ-
ent observed distributions. Anyway, we can safely
state that we have a maximum deviation between
the observations and the model distributions of
about a factor of 1.8 (being the observations lower
than the model).
The scaling law is weakly dependent on the den-
sity, so a change in the numerical values affects
the MPF in a negligible way. Notice also that
craters larger than about 0.1-0.2 km form in grav-
ity regime, hence the rock strength is not impor-
tant at those sizes. For smaller sizes, the shape of
the MPF may depend on the local strength. The
density and strength profiles adopted in this work
are meant to be average profiles for the Moon.
Local variations to the mean values may slightly
change the MPF at small crater sizes.
A more important parameter is the assumed depth
for the transition from cohesive to hard rock scal-
ing law. A variation in this value causes a shift of
6
the inflection point in the MPF within the range
1 < D < 10 km.
Neukum & Ivanov (1994) results within their er-
ror bars3.
4. A new chronology for the Moon
4.1. Moon cratering
The MPF can be used as a reference to develop
the lunar chronology. For this purpose, we used
literature crater counts from regions with known
radiometric ages as calibration regions. A list of
the regions used is reported in table 1.
In order to obtain the lunar chronology, we first
assumed that the MPF shape remained constant
over time. For each calibration region we deter-
mined the proportionality factor which gave the
best fit to the data by minimizing the χ2. Fi-
nally, we determined the crater cumulative num-
ber at 1 km (N1). The fitting procedure used is
particularly important, as it directly affects the
chronology. For this reason, we performed several
tests in order to assess its stability. A major as-
pect consisted in choosing the weights given to the
measured data points.
It is common practice in
crater counting to assign each data point an error
bar corresponding to the square root of the num-
ber of counts, under the assumption that the mea-
surements follow a Poisson distribution. However,
in the case in which significant systematic effects
are present, this procedure leads to underestimate
the total error. This problem is particularly evi-
dent when dealing with small craters that are at
the limit of the image resolution. Because of their
large number, the statistical error for these craters
is comparatively small. On the other hand, due to
the difficulty of positively identifying craters at the
limit of the resolution, these measurements could
be affected by a significant bias. This effect may
affect the N1 values up to a factor of 3-4 in some re-
gions and consequently have a non negligible effect
on the chronology. After several tests, we decided
to use the following weights:
t = N + κN 2
σ2
(10)
where κ was set to 0.5. In this way, the total er-
ror (σt) for points with very low statistical σ were
increased, while those with large statistical error
(i.e. low number of craters) were basically left un-
changed.
In order to validate this procedure, we applied
it to the old NPF, being able to reproduce the
In this section we present the results of the fits
applied to the regions used to calibrate the lunar
chronology (see tab. 1). The fitting procedure
was successful for most of the regions.
In sev-
eral cases a very good fit was achieved through
the measured range of the cumulative distribu-
tions,
indicating an overall agreement between
the shape of the MPF and the data. However,
some discrepancies occurred for the highlands (for
D > 100 km), Mare Crisum and Mare Tranquil-
litatis. In the case of the highlands (see fig. 5),
the discrepancy seems to suggest a different shape
of the impactor size distribution in the past, with
a larger fraction of medium-to-large impactors
(see section 5 for further details).
In the case
of Mare Crisum and Mare Tranquillitatis regional
phenomena can be responsible for the observed de-
viations. In particular, in both cases several con-
secutive magma flows have partially covered the
population of small craters, affecting the observed
cumulative distribution (Neukum & Horn 1976;
Boyce et al. 1977). Indeed the absolute radiomet-
ric ages derived from A11 samples (Mare Tran-
quillitatis), span from 3.5 to 3.85 Gyr, whereas
the ages of the Luna 24 basalts (Mare Crisium)
are clustered into at least three different peaks
at 2.5, 3.3, 3.6 Gyr, respectively (Birck & All`egre
1978; Stoffler & Ryder 2001; Fernandes & Burgess
2005). In presence of such surface phenomena, the
fit must be constrained on the basis of geological
considerations, as the automatic fit may be mis-
leading.
In particular, as formerly suggested by
Neukum & Horn (1976), the crater size-frequency
distribution derived from these regions leads to a
composite curve in which smaller craters reflect
the younger radiometric ages, while the larger
ones correspond to older ages. Thus, for Mare
Tranquillitatis we forced the MPF to overlap the
cumulative curve at 0.5 km for the age of 3.55 Gyr
(referred to as young), and at 0.9 km for the age
of 3.72 Gyr (referred to as old). Similar consider-
ations hold also for Mare Crisum where the MPF
3Here we used the old NPF instead of the new NPF because
even the latest chronology by Neukum et al. (2001) uses
the old NPF.
7
was fitted to craters with diameters in the range
1 to 2 km (details on the fit of the calibration
regions are reported in the online material).
Another effect influencing the accuracy of the
crater statistics is the possible presence of sec-
ondary craters. Even though authors publish-
ing crater counts pay particular attention to the
identification and rejection of secondary craters,
in some cases confusion with primary craters is
difficult to avoid. According to Ivanov (2006),
secondaries are likely negligible on young regions
because they did not have the time to accumu-
late in large numbers. On the other hand, older
regions, namely the highlands and maria, may
show secondary craters caused by subsequent
large impacts (Wilhelms 1976). Nevertheless,
Neukum & Ivanov (1994) concluded that the sec-
ondaries are likely not relevant. For the present
analysis, we assumed that the observed crater
cumulative distributions are resulting from the
primary flux of impactors.
4.2. Earth cratering
Cratering studies are much more difficult on the
Earth than on the Moon, due to strong resurfacing
processes and consequent alteration of the crater
morphologies. Nevertheless, Earth craters are im-
portant because they can be studied in detail and
precisely dated. Therefore they can be used to
further constrain the flux of impactors, in partic-
ular for young ages, that are scarcely represented
in lunar data. In this work two data sets of terres-
trial craters have been considered. The first one
consists of all large craters (D > 20 km) found
on the North American and Euroasiatic cratons,
as presented by Grieve & Dence (1979) (updated
with few new findings from the Earth Impact
database4). Following Grieve & Dence (1979), we
have adopted for these two cratons an age of
0.375 Gyr. Some sectors of both cratons were
probably exposed to the meteoroid flux already
from the early Ordovician (0.450 Gyr), whereas
others only from 0.300 Gyr ago (Grieve & Dence
1979). The choice of 0.375 Gyr is further sup-
ported by the age of the older craters that, with
only few exceptions, are around this value.
The second data set represents a subset of the
first one, including only craters with radiometric
4http://www.unb.ca/passc/ImpactDatabase/.
ages less then 0.120 Gyr. This choice has been
made following Grieve (1993), in order to over-
come biases due to erosion and burial of craters
on the Earth's surface. Both cumulative distri-
butions were fitted using our Earth MPF, and the
obtained N1 was rescaled to the lunar case by con-
sidering the gravitational focusing and the differ-
ent relationships between impactor and crater size
on the two bodies (see fig. 6). The resulting val-
ues for N1 are shown in tab. 1, with more details
supplied in the online material.
4.3. MPF chronology
The lunar chronology has been developed on
the basis of the relationship between the derived
N1 values and the absolute ages. However, this
task is not trivial since only 18 measurements are
available, 12 of which are older than 3 Gyr. More-
over, the region from 1 Gyr to 3 Gyr lacks com-
pletely in data points. Therefore a very accurate
chronology cannot be expected, in particular for
ages younger than 3 Gyr. In this work the radio-
metric ages proposed by Stoffler & Ryder (2001)
were used.
In order to describe the lunar chronology, Neukum
(1983) suggested the following approximation:
N1 = a(ebt − 1) + ct.
(11)
This function assumes a linear relationship be-
tween N1 and the age (t) for bt ≪ 1, while for
bt ≫ 1 N1 increases exponentially. Although this
is a rather simple description of the data, it accu-
rately fits the data points in the Neukum chronol-
ogy. We fitted the same function to our MPF-
based data points, with the results shown in fig.
7. The derived coefficient are: a = 1.23 × 10−15,
b = 7.85, c = 1.30 × 10−3. For comparison,
Neukum & Ivanov (1994) obtained: a = 5.44 ×
10−14, b = 6.93, c = 8.38 × 10−4, which are quite
similar to our result5. The Neukum & Ivanov
5As previously discussed, old regions with cumulative count-
ing extending to sub-km sizes may be affected by sec-
ondaries. These regions are: Descartes Formation, Mare
Imbrium, Mare Fecunditatis, Taurus Littrow Mare, Mare
Tranquillitatis. Therefore we also performed a best fit
excluding these regions. The resulting coefficients are
a = 3.07 × 10−15, b = 7.63, c = 1.33 × 10−3, and the
resulting chronology curve is basically overimposed to the
one obtained using all the regions. Therefore the secondary
craters -if presents in the cumulative distributions- do not
affect the chronology.
8
(1994) chronology curve is also plotted in fig. 7.
While the derived best fit is in overall agreement
with the nominal MPF-based data points, there
are some deviations that mainly concern young
data points, which lie systematically above the
best fit curve. It is remarkable that in our chronol-
ogy, the proposed equation fits well all the points
older than Copernicus (which was an outlier in
the chronology of Neukum & Ivanov 1994), while
it underestimates -by a factor of 2- all the points
younger than Copernicus (see fig. 7).
4.4. Non-constant flux in recent times?
As mentioned, the MPF-based chronology is
reasonably in agreement with that derived from
the NPF. The discrepancies observed for young
regions could be due to uncertainties involved in
our model.
In particular, a concern regards the
slope of the impactor cumulative size distribution
for small dimensions which is poorly constrained
by observations. Errors in the slope in this size
range may result in inaccurate estimates of N1 for
the young lunar regions. The slope of the NEO
model for impactor diameters smaller than 0.1 km
(which roughly corresponds to 1 km crater size) is
-2.6, which becomes a slope of -3.1 in the MPF.
This slope is very similar to that of the NPF for
the same size range, which is equal to -3.2. This
slight difference in slope can account for half of
the observed difference between our and Neukum's
chronology. We underline that such slope varia-
tions may affect the Cone, Tycho and North ray
craters, but not the terrestrial crater counts since
they have D > 20 km. Therefore it is remark-
able that terrestrial craters lie on the same linear
trend of the young lunar data points (fig. 7, left
panel). The NEO population model adopted here
is in overall agreement with the available observed
data (see fig. 16 of Bottke et al. 2005b). For small
dimensions, however, the real NEOs size distri-
bution is poorly known. The only observational
constraints are the small fireballs (Halliday et al.
1996) and bolides (Brown et al. 2002) detected in
the Earth's atmosphere. A careful check at fig.
16 of Bottke et al. (2005b) shows that the NEO
population model slightly overestimates -by a fac-
tor of about 1.5- the predicted number of NEOs
from bolide data. This occurs for the impactor size
range responsible for the cratering observed for the
Cone, North Ray and Tycho craters. Based on the
9
data presently available, it is very difficult to un-
derstand whether this difference is real or rather
due to the involved uncertainties. Nevertheless,
if we maintain the present number of NEOs at
d = 1 km, we obtain that the N1 values corre-
sponding to the younger lunar regions would be
further increased.
Other possible sources of uncertainties may arise
from the scaling law. Nevertheless, variations in
the parameters produce a vertical shift of the MPF
and not a change of the shape, therefore this shift
alone cannot affect the determination of N1 values.
The only way to obtain a change in the N1 values
with respect to the nominal model, is to introduce
a slope variation in the MPF for D < 1 km. This
would be the case, if the inflection point would
slide below D ∼ 1 km. This would imply a very
low thickness for the fractured layers which seems
improbable, according to our understanding of the
lunar crust. Nevertheless, if this would be the
case, N1 for the young lunar regions would be fur-
ther increased.
It is also possible that the MPF-based N1 values
are accurate, and the simple relationship proposed
by Neukum (1983); Neukum & Ivanov (1994) is
not adequate to describe the data. This would be
the case, for example, if the impactor flux had not
been constant during the past ∼3 Gyr. Due to the
lack of data points in the range 1-3 Gyr, we are not
able to study in detail possible variations in the
flux. However, fig. 7 seems to suggest that a recent
phase (age < 0.4 Gyr) characterized by a constant
impactor flux was preceded by a phase (between
0.8 and 3 Gyr) with a lower and nearly constant
impact rate. A single event placed around 0.4 Gyr
ago and lasting for 0.2-0.3 Gyr that would increase
the impactor flux by a factor 2-3 would explain
the observed data. Such a recent change in im-
pactor flux would also have affected the older re-
gions, but here the relative increase in the number
of craters would be very low compared to the accu-
mulation of craters over 3 Gyr or more, hence this
effect would be negligible. This scenario is quali-
tatively in agreement with the recent suggestion of
an increase in the number of impactors in the in-
ner Solar System due to the formation of dynam-
ical families -as a results of catastrophic disrup-
tion of asteroids- in the main belt (Nesvorn´y et al.
2007; Bottke et al. 2007). In particular, the Bap-
tistina and Flora families are estimated to have
formed about 140 Myr and 500 Myr ago, respec-
tively. Their proximity to the resonances led to
an increase in the member of the NEO population,
starting few tens of Myr after their formation, as a
consequence of the Yarkowsky-effect-driven decay
of the semimajor axis. In the case of Baptistina
the maximum flux was reached about 60 Myr af-
ter its formation (i.e. 80 Myr ago). Therefore this
flux may have affected the Cone, North Ray and
Tycho craters counting6, and also the young ter-
restrial craters.
Concerning the formation of Flora, similar consid-
erations may hold. In addition, the Flora family
has been connected to the spike in the infall of L-
chondrite meteorites that occurred about 470 Myr
ago (Bogard 1995). A similar spike has also been
detected in the lunar glass spherules and -to a
lesser extent- in lunar impact clasts (Culler et al.
2000; Cohen et al. 2005). These spikes may have
affected mostly the terrestrial cratering and the
Copernicus counting. Although such enhance-
ment of the flux was probably formed by sub-
km impactors (Hartmann et al. 2007), the spike
recorded in the lunar melt clasts implies also the
existence of a number of km-sized impactors. This
enhancement in the km-sized impactors may ex-
plain why the Earth cratering for D > 20 km
seems not to be depleted by erosion processes (fig.
7, left panel), as expected (Grieve 1993).
4.5. On the early flux
One of the major -and mostly debated- open
issues is related to the cratering rate in the early
times after the formation of the Moon. Argu-
ments have been proposed in favor of a rapid
and smooth decrease in the impact rate from
the formation of the Moon till about 3.5 Gyr
ago (Hartmann et al. 1981; Neukum 1983), fol-
lowed by a constant impact rate (Neukum 1983;
Neukum & Ivanov 1994). On the other hand,
studies of impact melts led some researchers (e.g.
Ryder 1990; Stoffler & Ryder 2001; Cohen et al.
2000) to support the idea of an intense lunar bom-
bardment about 3.9 Gyr ago which lasted some
100-200 Myr. From a dynamical point of view,
such short phase of intense bombardment has been
6See the discussion in Bottke et al. (2007) about the prob-
ability of Tycho itself being formed by an impact with a
Baptistina fragment.
recently explained in the more general context of
the early stages of the evolution of the Solar Sys-
tem (Gomes et al. 2005). This scenario has also
found some observational evidence in the work of
Strom et al. (2005).
Despite these works, the LHB hypothesis has not
found an unanimous consensus yet. For instance,
Hartmann et al. (2007) raised doubts about the
interpretation of lunar melts and glasses data;
while Neukum & Ivanov (1994) argued against the
LHB on the basis of the smooth behavior of their
lunar chronology curve.
If the intense LHB did take place, it should have
left some traces in the chronology curve, which
therefore could be used to obtain constraints on
the early stages of the cratering. Unfortunately,
there are only 5 measured regions having ages
older than 3.85 Gyr, and therefore it is not easy to
draw firm conclusions. One of the most important
regions is the Nectaris Basin. Neukum & Ivanov
(1994) used a radiometric age of 4.1 Gry. The re-
sulting best fit for the lunar chronology curve is
therefore very close to all the old regions (within
errors) and has a smooth behavior in early times,
reflecting a possible smooth decay in the impact
rate (see fig. 7). However, using the new esti-
mate of 3.92 Gyr for Nectaris Basin proposed by
Stoffler & Ryder (2001) this point moves consider-
ably far from the best fit (this is the age we used
in our chronology). This results in a change of the
slope of the chronology curve, which in turn is in
favor of the LHB. We underline that in the present
analysis we used the present NEO size distribu-
tion. However, it is very likely that at the time
of the LHB the impactor size distribution had a
different shape. At the present stage of knowledge
a firm conclusion from the chronology curve can-
not be obtained. Nevertheless, we will add few
considerations on this important point in the next
section.
5. A time-dependent MPF
In this section we develop a time-dependent
MPF. In doing so, we clearly show the versatility
of our approach, as opposed to the conventional
methods (e.g.
the NPF), where the production
function is generally assumed to be constant over
time. Modeling a time-dependent MPF is quite
complex, and in this section we will only pose the
basis of the problem, deferring thorough investiga-
10
tions to a further analysis. In general terms, there
are two possible reasons for such time-dependent
behavior: The first one is related to a non sta-
tionary flux7, φ(d, v, t); the second one to crater
erasing processes.
A motivation for studying the non stationary flux
is provided by the dynamical processes responsi-
ble for the flux. In early times (just after the for-
mation of the terrestrial planets and the Moon,
some 4.5 Gyr ago) the inner Solar System was
still populated by the leftover planetesimals, which
were rapidly (10-100 Myr) cleared-up by the latest
stages of the accretion process. A second source
of impactors was represented by the highly per-
turbed bodies (e.g.
through the sweeping reso-
nance mechanism) due to the formation and mi-
gration of Jupiter. At the present time, the flux
is mainly sustained by the slow decay of MBAs
into the resonances. All these three stages have
their own characteristic φ(d, v). For sake of sim-
plicity, here we focus on the last two mechanisms.
The second process is size dependent, while the
first one is not, therefore the shape of the im-
pactor size distribution changes over time. On the
other hand, the impactor velocity distribution is
expected to remain approximately constant over
time, as in both cases the transport mechanism
to the NEO space is the same. Therefore we may
write φ(d, v, t) = h(d, t)f (d, v).
The crater erasing on large bodies depends mainly
on two major effects: the superposition of craters
and the local jolting (O'Brien et al. 2006). The
first one basically accounts for the overlapping of
craters as the surface crater density increases. The
second one considers the erasing of craters due to
local seismic jolting triggered by the formation of
large craters. Both effects can be modeled by con-
sidering a variation in the number of craters in a
given size bin (O'Brien et al. 2006). Let E be the
ratio between the actual number of craters consid-
ering the erasing and the total number neglecting
the erasing, we can write:
MPF(D, t) = Z ∞
D
Φ( D, t)E( D, t) d D
(12)
7We stress once again that in our approach what is relevant
is the shape of the impactor size and velocity distributions,
since the absolute calibration of the model is done by the
reference lunar regions.
where setting t = 0 and E = 1 reduces to
the MPF used in previous sections (see appendix
for further details). Here we limit ourselves to
considering these effects separately. Let us con-
sider first the crater erasing mechanism, assuming
φ ≡ φ(d, v, 0) as in previous sections. The re-
sults are shown in fig. 8. It is clear that as the
age increases the MPF becomes flatter for small
dimensions. This is the well-known saturation
process of heavily cratered surfaces. This affects
the derived N1 values. To study possible varia-
tions in the rate of impacts over time, we adopted
the following procedure: first the cumulative dis-
tributions for the calibration regions were fitted
with the MPF corrected for the erasing, then the
obtained N1 was rescaled to the total number of
craters that occurred in the region (i.e. the actual
number of impactors). In fig. 9 the effects of the
erasing on the chronology are shown. Although
the erasing considerably changes the MPF for the
old ages (see fig. 8), its effects on the chronology
are not important.
Let us now deal with changes in the impactor size
distribution and no erasing (E ≡ 1). Following
the previous discussion, we assumed that before
the end of the LHB (∼ 3.8 Gyr ago) the size dis-
tribution h(d) resembled closely that of the main
belt8 (hm), while afterwards it becomes similar
to the present one, that is hn. This assumption
is also encouraged by the observation that hm
provides a much better fit to the crater distribu-
tion of the highlands and Nectaris Basin than hn
does (see online material). With respect to the
previous analysis (see fig. 7), the only difference
regards two data points, namely the highlands
and Nectaris Basin. For both regions we obtain
a lower N1 value, with the reduction being more
pronounced for the highlands. This results in a
net change in the slope of the chronology in the
two oldest points, suggesting a change in the rate
of impactors. Therefore, if the MBA size distri-
bution is used for earlier times on the basis of the
improved fit with both the highlands and Nectaris
Basin, the chronology plot would suggest the ex-
istence of the LHB. These conclusions are weakly
affected by the uncertainties in the model (size
distribution, scaling law etc.) because the MPF is
8For sake of simplicity, we used the present MBA size dis-
tribution although it may have had a different shape in the
past (Bottke et al. 2005a,b).
11
well defined for the large crater sizes relevant in
this context.
6. Mercury's chronology
The impactor size and velocity distributions for
Mercury have been computed in a similar way as
done for the Moon (figs 10, 11). We assumed den-
sity and strength profiles similar to those adopted
for the Moon. The resulting MPF is shown in
fig. 12 (right panel), along with the overplot of
the NPF for Mercury (Neukum et al. 2001b). The
two production functions nicely overlap at the ex-
tremes, that is for D < 0.3 km and D > 40 km. In-
side this range, however, the two curves are quite
different, with a maximum deviation of nearly a
factor of 5 at D ∼ 3 km (fig. 12, left panel). The
exact behavior of the MPF in this range depends
on the assumptions on the transition in the scaling
law between cohesive soil and rock (see discussion
in sect. 2.2). Nevertheless, whatever reasonable
choice is made for the scaling law parameters, our
MPF is remarkably different from the NPF in this
range.
As an example, we report here our age determi-
nation of the Chekhov basin. Crater counting
was performed on the total ejecta blanket, over
an area exceeding 4 · 104 km2. According to the
analysis of Neukum et al. (2001b), Chekhov basin
has a derived age of 4.05 Gyr, which makes it
one of the oldest regions on Mercury. In fig. 13
we report our MPF-based age estimate, obtained
considering different scenarios, namely using the
NEO and MBA size distributions. In the nominal
case (i.e. MPF=MPF(D, 0); E = 1), we obtain
an age of 3.94 Gyr, which is slightly younger than
that found by Neukum et al. (2001b). The corre-
sponding χ2 of the best fit is 0.8. Note that, in
comparison to the NPF (χ2=1.5), the MPF more
closely fits the observed crater cumulative distri-
bution. When considering also the crater erasing,
the age becomes 4.05 Gyr (χ2=0.6), and the shape
of the MPF is basically unchanged.
In the case
of the MBA size distribution and no erasing, we
obtain 3.97 Gyr (χ2=0.6). Notice that in the case
of MBA, if we include crater erasing the MPF
reaches the saturation prior to achieving the best
fit, therefore it is not possible to derive a precise
age estimate.
In conclusion, the MPF shows a
better fit to the data points than the NPF. How-
12
ever, due to the measurement errors and to the
limited crater size range observed, it is not pos-
sible to conclude whether the NEO or the MBA
size distribution is more suitable for fitting the
Chekhov basin.
7. Conclusions
In this paper a new model for the chronology
of planetary surfaces in the inner Solar System
has been introduced and applied to the Moon and
Mercury. Concerning the Moon, the main findings
are the following:
• The nominal MPF differs from the NPF for
less than a factor of 2 throughout the ob-
served range of dimensions. A similar degree
of agreement is obtained for the chronology;
• Despite the good agreement between MPF
and NPF, there is a systematic misfit of
the assumed linear branch of the chronology
function with the data points for ages more
recent than about 0.4 Gyr.
It is remark-
able that the Earth craters (which have D >
20 km) are aligned with the Tycho, Cone
and North Ray craters (having D < 0.4 km).
Although the misfit may be partially due to
the uncertainties present in the model, other
causes, such as for instance a non-constant
flux in recent times due to the formation
of MB families (like Baptistina and Flora)
might be possible;
• The MPF was computed for two differ-
ent impactor size distributions, namely the
present NEO and MBA populations.
In
this way we were able to study whether the
cratering production function changed over
time. Both the NEO and MBA-based MPFs
are able to fit the cumulative distributions
having ages < 3.9 Gyr. This is mainly due
to the fact that such regions have craters
smaller than ∼ 10 km, where the MPFs are
very similar. On the other hand, the old-
est regions (highlands 4.35 Gyr; Nectaris
Basin 3.92 Gyr) are best fitted using the
MBA size distribution. This suggests that
the impactor size distribution around 4 Gyr
ago resembled the present MBA distribu-
tion, while today it is closer to the present
NEO size distribution. Alternatively, it is
also possible that comets had in the past a
non negligible contribution. This point de-
serves a specific analysis and will be subject
of a future work;
Concerning Mercury, the MPF differs from the
NPF in the size range from 0.3 km to 40 km. The
maximum difference -of about a factor of 5- oc-
curs at ∼ 3 km. This discrepancy is therefore
potentially important for dating purposes, in par-
ticular when measurements for small craters will
become available (e.g.
from MESSENGER and
BepiColombo). Outside this range, the NPF and
MPF are very similar. We show the results of our
model for the case of the Chekhov basin. In the
nominal case (MPF(D, 0), E = 1) we obtained the
age of 3.95 Gyr, which compares to the 4.05 Gyr
of Neukum et al. (2001b).
We are grateful to A. Morbidelli for helpful dis-
cussions and for providing us with the results of
the orbital integrations. We are also grateful to
R. Wagner for providing us with the original lunar
cumulative distributions. Thanks to G. Neukum
for helpful discussion and suggestions. We also
thank W. Bottke for providing us with the size
distribution of NEOs and MBAs. Finally, thanks
to D. De Niem and G.F. Gronchi for helpful dis-
cussions.
13
A. The Model Production Function
The cumulative distribution of craters (the so-called production function) can be evaluated starting from
the differential distribution of the incoming flux, namely φ(d, v, t). Here we consider the general problem of a
non stationary flux, therefore the time dependence of the function φ. Each impact produces a crater, with a
size specified by the scaling law, D = S(d, v, ~p). Let Φ(D) be the differential distribution of craters, namely
the number of craters per unit of crater size per unit surface. Such distribution is obtained by integrating the
differential distribution of impactors over the one-dimensional domain specified by the scaling law, namely
γD : S(d, v, ~p) = D. Let σ be a parameter along the curve γD, we finally have:
dσ
(A1)
Φ(D, t) = Z φ(γD(σ), t)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
dγD
dσ
(σ)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)
In the presence of erasing processes, the actual number of measured craters in any given size bin dD is
not equal to the number of impactors. Let E(D, t) be the ratio of the final number of crater erasing included,
with the total number (i.e. erasing excluded), we finally obtain:
MPF(D, t) = Z ∞
D
Φ( D, t)E( D, t) d D
Following the work of O'Brien et al. (2006), the function E can be expressed as
E(D, t) =
1 − e−A
A
(A2)
(A3)
where A is the total effective area of craters accumulated per unit area. Therefore A depends on the
age since older regions have a large number of craters accumulated and hence a larger A with respect to
younger regions. A also depends on crater diameter since small craters are more frequent than larger ones.
As a results, the counting of small craters on old regions are strongly affected by the erasing (see fig. 8). In
calculating A, we considered the superposition of craters and the local jolting, and neglected global jolting
and cumulative effects of seismic shaking, since they are not relevant for large bodies like the terrestrial
planets. The parameters involved in the estimate of A have been derived from O'Brien et al. (2006). The
implementation of the erasing processes in our code has been tested accurately using asteroid crater counting
(Gaspra, Ida, etc). Our results agree with those published. Nevertheless, the purpose of the present analysis
is simply to show the first-order effects of the erasing, and a detailed evaluation of the erasing on planetary-
sized bodies is left for further analysis.
14
Description of the lunar and terrestrial calibration regions used in this work.
Table 1
Region
N1 (NEO)†
(km−2)
N1 (MBA)†
(km−2)
Age‡
(Gyr)
Highlandsa
Nectaris Basin
Descartes Formationb,s
Imbrium Apenninesc
Fra Mauro Formation
Mare Tranquillitatis (old)d,s
Taurus Littrow Mares
Mare Tranquillitatis (young)d,s
Mare Fecunditatiss
Mare Imbriums
7.851 · 10−1
1.327 · 10−1
2.490 · 10−2
1.968 · 10−2
2.595 · 10−2
1.836 · 10−2
1.579 · 10−2
9.300 · 10−3
3.234 · 10−3
5.468 · 10−3
Mare Crisium 2.335 · 10−3
Oceanus Procellarum 3.683 · 10−3
1.321 · 10−3
1.348 · 10−3
1.267 · 10−3
3.835 · 10−4
3.391 · 10−4
1.644 · 10−4
1.389 · 10−4
6.970 · 10−5
Copernicus Crater (cont. ejecta)e
Copernicus Crater (crater floor)e
Terrestrial Phanerozoic cratersf
Terrestrial Phanerozoic craters (young)g
Tycho crater (cont. ejecta)h
Tycho crater (cont. ejecta)i
North Ray craterl
Cone Craterl
2.018 · 10−1
6.648 · 10−2
2.509 · 10−2
1.931 · 10−2
2.672 · 10−2
1.832 · 10−2
1.585 · 10−2
9.357 · 10−3
3.257 · 10−3
5.526 · 10−3
2.377 · 10−3
3.695 · 10−3
1.337 · 10−3
1.343 · 10−3
7.655 · 10−4
2.195 · 10−4
3.401 · 10−4
1.712 · 10−4
1.421 · 10−4
7.131 · 10−5
4.35
3.92
3.92
3.85
3.85
3.80
3.70
3.58
3.41
3.30
3.22
3.15
0.80
0.80
0.375
0.120
0.109
0.109
0.053
0.025
Note. -- Where not explicitly quoted, the data are from Neukum (1983) and refer-
ences therein. The plots showing the fits of the calibration regions are reported in the
online material.
†The derived N1 values both using the NEO and MBA size distributions.
‡Radiometric ages for the lunar regions are from table VI of Stoffler & Ryder (2001).
The ages of the terrestrial craters are from Grieve & Dence (1979).
aAll craters and basins.
bThe counting for Descartes Formation is limited to dimensions smaller than 1.2 km,
while that used in Neukum (1983) extends to larger sizes.
cThe crater counting for Imbrium Apennines (used in this work) is a sub sample of
the crater counting from Imbrium Basin, used by Neukum (1983).
dOriginal data from Greeley & Gault (1970).
eFor the Copernicus crater we have two distinct measurements corresponding to the
crater floor and the continuous ejecta blanket. In fig. 31 of Neukum (1983), these sets
were probably merged into a single distribution. We decided to keep these two sets
separate, and to provide individual fits (see fig. 7 and the online material).
f Terrestrial
and
updated
are
to
Phanerozoic
craters
according
from Grieve & Dence
(1979),
Database
(http://www.unb.ca/passc/ImpactDatabase/). Only large craters (D > 20 km)
on the North American and Euroasian cratons have been considered. For sake of
completeness we report the list of craters used.
In the North American craton:
Beaverhead, Carswell, Charlevoix, Clearwater East, Clearwater West, Haughton,
Manicouagan, Manson, Mistastin, Presquile, Saint Martin, Slate Islands, Steen River,
Sudbury.
In the Eurasian craton: Boltysh, Kamensk, Keurusselka, Lappajarvi,
Puchezh-Katunki, Siljan.
derived
the
Earth
Impact
gOnly large (D > 20 km) and young (< 120 Myr) terrestrial craters. They are:
Carswell, Haughton, Manson, Mistastin, Steen River, Boltysh, Kamensk, Lappajarvi.
h,iTwo distinct measurements were available for the Tycho crater. Both correspond
to the continuous ejecta: in (h) small craters were counted, while in (i) large craters
were counted. In fig. 31 of Neukum (1983) these files were merged into a single dis-
tribution. We decided to keep these two sets separate, and to provide individual fits
(see fig. 7). Notice that Neukum (1983), also reports 4 measured points in the range
0.5-1.1 km that were not available to us.
lOriginal data from Moore et al. (1980).
15
sThese are relatively old regions having crater counting extending to small sizes
(< 1 km) and thus potentially affected by secondary craters. In order to check whether
these regions affect the final MPF-based chronology, we also performed a best fit ex-
cluding these points. The results of the fit is basically unchanged (see text for further
details.)
16
Moon (S)
Moon (L)
Earth (S)
Earth (L)
0.12
0.1
0.08
0.06
0.04
0.02
y
t
i
l
i
b
a
b
o
r
P
0
0
10
20
30
Impact velocity (km/s)
40
50
Fig. 1. -- Distributions of impact velocities on the Moon and on the Earth, for two different impactor sizes.
The smallest (S) and largest (L) impactor sizes used in these simulations are 0.1 m and 72 km, respectively.
The average impact velocities are 18.6 and 20.0 km/s, respectively. Notice that these average impact velocities
are slightly lower then those used by Neukum & Ivanov (1994), after applying the correction for the average
impacting angle of π/4. The distributions are quite spread and therefore a large number of impacts occur
at velocities considerably different from the average, thus affecting the final crater distribution (see text).
17
)
1
-
2
-
r
y
m
k
(
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
0.0001
1e-06
1e-08
1e-10
1e-12
1e-14
1e-16
1e-18
1e-20
0.0001 0.001
Moon
Earth
MBA
0.01
0.1
Impactor diameter (km)
1
10
100
Fig. 2. -- Model cumulative size distributions of impactors on the Moon and on the Earth (after Bottke et al.
2002). An arbitrarily rescaled MBA cumulative distribution (after Bottke et al. 2005a) is also shown for
comparison. The main belt distribution shows an S-shaped feature in the 0.2-2 km size range. Also the
NEO distribution shows a qualitatively similar feature. The two distributions, however, have quantitatively
different shapes, due to selection mechanisms taking place during the dynamical transport from the main
belt to the near-Earth space.
18
Moon
Earth
200
175
150
125
100
75
50
25
)
m
k
(
h
t
p
e
D
Moon
Earth
80
70
60
50
40
30
20
10
)
m
k
(
r
e
t
e
m
a
i
d
r
o
t
c
a
p
m
I
0
2
2.5
3
Target density (g/cm3)
3.5
0
2
2.5
3
3.5
Target density (g/cm3)
lunar and terrestrial density
Fig. 3. -- Assumed density profiles for the Moon and the Earth. Left panel:
profiles vs depth. Right panel: average density vs impactor size. The average density has been obtained
by averaging the density up to a depth of 10× the impactor radius. A similar consideration holds for the
strength (see text for details).
1e-06
1e-08
)
1
-
NPF new
NPF old
MPF (H&H)
MPF (I&S)
MPF (S&S)
2
-
r
y
m
k
(
s
r
e
t
a
r
c
f
o
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
1e-10
1e-12
1e-14
1e-16
1e-18
0.01
0.1
1
10
Crater diameter (km)
100
Fig. 4. -- Comparison of the MPF obtained with different scaling laws, indicated by H&H, I&S, S&S (see
section 2.2 for details). The old and new NPFs are also shown.
19
0.0001
1e-05
)
2
-
m
k
(
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
1e-06
Highlands
MPF
NPF old
1e-07
10
100
Crater diameter (km)
1000
Fig. 5. -- The plot shows the best fit of the crater-size cumulative distribution for the lunar highlands with
both the MPF and the NPF.
3
2
1
o
i
t
a
r
F
P
M
0
0.01
Earth/Moon
Moon
Earth
)
1
-
2
-
r
y
m
k
(
s
r
e
t
a
r
c
f
o
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
100
0.01
0.1
1
10
Crater diameter (km)
0.1
1
10
Crater diameter (km)
1e-06
1e-09
1e-12
1e-15
1e-18
100
Fig. 6. -- Detailed comparison of the lunar and terrestrial MPFs (using the H&H scaling law).
20
Neukum & Ivanov 1994
MPF chronology
Neukum & Ivanov 1994
MPF chronology
1e+00
1e-01
1e-02
)
2
-
m
k
(
1
N
1e-03
Copernicus
Terrestrial
Tycho
North Ray
Cone
1e-04
1e-05
0.01
0.1
Age (Gyr)
1
0
1
2
Age (Gyr)
3
4
1e+00
1e-01
1e-02
1e-03
1e-04
1e-05
Fig. 7. -- The plot shows the N1 values for the calibration regions obtained using the MPF and those obtained
by Neukum & Ivanov (1994). The solid curves, namely the chronology curves, represent the best fit of the
two sets of data points. The shadowed regions encompass a factor of ±2 around the MPF chronology curve.
21
0.5 Gyr
1.5 Gyr
2.5 Gyr
3.5 Gyr
4.0 Gyr
4.5 Gyr
0.5 Gyr with CE
3.5 Gyr with CE
4.0 Gyr with CE
4.5 Gyr with CE
1e+04
1e+03
1e+02
1e+01
1e+00
1e-01
1e-02
1e-03
1e-04
1e-05
1e-06
1e-07
)
2
-
m
k
(
s
r
e
t
a
r
c
f
o
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
1e-08
0.01
0.1
1
10
Crater diameter (km)
100
Fig. 8. -- Moon isochrones obtained with MPF(D, 0), and including the crater-erasing (CE) mechanism.
The effect of longitude variations is also shown by the shadowed area for the 4.5 Gyr curve (with no erasing).
The expected longitudinal asymmetry in the impactor flux has been modeled using eq. 1 of Morota et al.
(2005). The impact of such analysis on the N1 values is negligible, therefore the chronology is not affected.
22
Neukum & Ivanov 1994
MPF chronology
MPF (NEO+MBA)
MPF with CE
Nectaris Basin
3
Age (Gyr)
4
1
0.1
0.01
0.001
Fig. 9. -- In this figure the effects of using the size distribution of MBAs and the crater-erasing (CE)
mechanism are shown for the old regions alone, since the young ones are not affected. The vertical strip
corresponds approximately the LHB event. Notice that for Nectaris basin the age from Stoffler & Ryder
(2001) adopted in this work differs from the one in Neukum & Ivanov (1994). Moreover, the age of the
highlands is known with large uncertainty and may vary in the range 4.2-4.4 Gyr. Here we adopted the
values used by Neukum & Ivanov (1994).
23
)
1
-
2
-
r
y
m
k
(
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
0.0001
1e-06
1e-08
1e-10
1e-12
1e-14
1e-16
1e-18
1e-20
0.0001 0.001
Moon
Earth
Mercury
MBA
0.01
0.1
Impactor diameter (km)
1
10
100
Fig. 10. -- The plot shows the model cumulative distribution of impactors for Mercury, in comparison with
that of the Earth-Moon system. The MBA size distribution is also shown.
0.12
0.1
0.08
0.06
0.04
0.02
y
t
i
l
i
b
a
b
o
r
P
0
0
10
20
30
Moon (S)
Moon (L)
Earth (S)
Earth (L)
Mercury (S)
Mercury (L)
60
70
80
90
40
50
Impact velocity (km/s)
Fig. 11. -- Impactor velocity distribution for Mercury, in comparison with the Earth-Moon system (see also
fig. 1).
24
5
4
3
2
1
/
F
P
N
F
P
M
0
0.01
)
1
-
2
-
r
y
m
k
(
s
r
e
t
a
r
c
f
o
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
100
0.01
0.1
1
10
Crater diameter (km)
NPF for Mercury
MPF
1e-06
1e-08
1e-10
1e-12
1e-14
1e-16
1e-18
100
0.1
1
10
Crater diameter (km)
Fig. 12. -- MPF for Mercury, in comparison with the NPF. The two production functions are in disagreement
in the range from about 0.1 km to 40 km.
Chekhov Basin (total ejecta)
NPF (4.05 Gyr)
MPF (3.94 Gyr)
MPF (with CE, 4.05 Gyr)
MPF (with MBA, 3.97 Gyr)
0.001
0.0001
)
2
-
m
k
(
r
e
b
m
u
n
e
v
i
t
a
l
u
m
u
C
1e-05
10
Crater diameter (km)
Fig. 13. -- MPF best fit of Chekhov basin and derived age. We also present the best fit using the MBA size
distribution and crater erasing (see section 5).
25
REFERENCES
Anderson, D. L. 2007, New Theory of the Earth.
Cambridge University press, 380 pp.
Asphaug, E., Moore, J. M., Morrison, D., Benz,
W., Nolan, M. C., & Sullivan, R. J. 1996,
Icarus, 120, 158
Birck, J. L., & All`egre, C. J. 1978, Physics of the
Earth and Planetary Interiors, 16, 10
Bogard, D. 1995, Meteoritics, 30, 244
Bottke, W. F., Vokrouhlick´y, D., & Nesvorn´y, D.
2007, Nature, 449, 48
Bottke, W. F., Durda, D. D., Nesvorn´y, D.,
Jedicke, R., Morbidelli, A., Vokrouhlick´y, D.,
& Levison, H. 2005a, Icarus, 175, 111
Bottke, W. F., Durda, D. D., Nesvorn´y, D.,
Jedicke, R., Morbidelli, A., Vokrouhlick´y, D.,
& Levison, H. F. 2005b, Icarus, 179, 63
Bottke, W. F., Morbidelli, A., Jedicke, R., Petit,
J.-M., Levison, H. F., Michel, P., & Metcalfe,
T. S. 2002, Icarus, 156, 399
Bottke, W. F., Jedicke, R., Morbidelli, A., Petit,
J.-M., & Gladman, B. 2000, Science, 288, 2190
Brown, P., Spalding, R. E., ReVelle, D. O., Tagli-
aferri, E., & Worden, S. P. 2002, Nature, 420,
294
Boyce, J. M., Schaber, G. G., & Dial, A. L., Jr.
1977, Nature, 265, 38
French, B. M. 1998. Traces of Catastrophe: A
Handbook of Shock-Metamorphic Effects in
Terrestrial Meteorite Impact Structures. LPI
Contribution No. 954, Lunar and Planetary In-
stitute, Houston. 120 pp.
Glass, B. P. 1990, Tectonophysics, 171, 393
Gomes, R., Levison, H. F., Tsiganis, K., & Mor-
bidelli, A. 2005, Nature, 435, 466
Greeley, R., & Gault, D. E. 1970, Moon, 2, 10
Grieve, R. A. F., & Dence, M. R. 1979, Icarus, 38,
230
Grieve, R. A. F. 1993, Vistas in Astronomy, 36,
203
Halliday, I., Griffin, A. A., & Blackwell, A. T.
1996, Meteoritics and Planetary Science, 31,
185
Hartmann, W.K., Strom, R., Weidenschilling, S.,
Blasius, K., Woronow, A., Dence, M., Grieve,
R., Diaz, J., Chapman, C., Shoemaker, E.,
Jones K., 1981. Chronology of planetary vol-
canism by comparative studies of planetary cra-
tering. In Basaltic Volcanism on the Terres-
trial Planets, pp. 1050-1129. Basaltic Volcanism
Study Project.
Hartmann, W. K., Quantin, C., & Mangold, N.
2007, Icarus, 186, 11
Holsapple, K. A., & Housen, K. R. 2007, Icarus,
191, 586
Cohen, B. A., Swindle, T. D., & Kring, D. A. 2000,
Science, 290, 1754
Koeberl, C., Poag, C. W., Reimold, W. U., &
Brandt, D. 1996, Science, 271, 1263
Cohen, B. A., Swindle, T. D., & Kring, D. A. 2005,
Meteoritics and Planetary Science, 40, 755
Culler, T. S., Becker, T. A., Muller, R. A., &
Renne, P. R. 2000, Science, 287, 1785
Farinella, P., Vokrouhlicky, D., & Hartmann,
W. K. 1998, Icarus, 132, 378
Fernandes, V. A., & Burgess, R.
2005,
Geochim. Cosmochim. Acta, 69, 4919
Ivanov, B. A. 2001, Space Science Reviews, 96, 87
Ivanov, B. A. 2006, Icarus, 183, 504
Ivanov, B. A., Neukum, G., & Wagner, R. 2001,
Astrophysics and Space Science Library, 261, 1
Ivanov, B. A., Neukum, G., Bottke, W. F., Jr., &
Hartmann, W. K. 2002, Asteroids III, 89
Marchi, S., Morbidelli, A., & Cremonese, G. 2005,
A&A, 431, 1123
Moore, H. J., Boyce, J. M., & Hahn, D. A. 1980,
Moon and Planets, 23, 231
26
Morbidelli, A., Bottke, W. F., Jr., Froeschl´e, C.,
Wilhelms, D. E. 1976, Lunar and Planetary Sci-
& Michel, P. 2002, Asteroids III, 409
ence Conference, 7, 2883
Morbidelli, A., & Gladman, B. 1998, Meteoritics
and Planetary Science, 33, 999
Morota, T., Ukai, T., & Furumoto, M. 2005,
Icarus, 173, 322
Nesvorn´y, D., Vokrouhlick´y, D., Bottke, W. F.,
Gladman, B., Haggstrom, T. 2007, Icarus, 188,
400
Neukum, G., & Ivanov, B. A. 1994, Hazards Due
to Comets and Asteroids, 359
Neukum, G., PhD Thesis, Meteoritenbombarde-
ment und Datierung Planetarer Oberflaechen,
Munich, Feb. 1983 p 1-186
Neukum, G., Ivanov, B. A., & Hartmann, W. K.
2001, Space Science Reviews, 96, 55
Neukum, G., & Horn, P. 1976, Moon, 15, 205
Neukum, G., Oberst, J., Hoffmann, H., Wagner,
R., & Ivanov, B. A. 2001b, Planet. Space Sci.,
49, 1507
O'Brien, D. P., Greenberg, R., & Richardson, J. E.
2006, Icarus, 183, 79
Pike, R. J. 1980, Icarus, 43, 1
Schmidt, R. M., & Housen, K. R. 1987, Interna-
tional Journal of Impact Engineering, 5, 543
Shoemaker, E.M., Wolfe, R.F., Shoemaker, C.S.,
1990. Asteroid and comet flux in the neighbor-
hood of Earth. Geological Society of America
Special Paper 247
Stoffler, D., & Ryder, G. 2001, Space Science Re-
views, 96, 9
Strom, R. G., Malhotra, R., Ito, T., Yoshida, F.,
& Kring, D. A. 2005, Science, 309, 1847
Stuart, J. S., & Binzel, R. P. 2004, Icarus, 170,
295
Ryder, G. 1990, EOS Transactions, 71, 313
Werner, S. C., Harris, A. W., Neukum, G., &
Ivanov, B. A. 2002, Icarus, 156, 287
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
27
|
1912.08994 | 1 | 1912 | 2019-12-19T02:44:14 | Crater-ray formation through mutual collisions of hypervelocity-impact induced ejecta particles | [
"astro-ph.EP"
] | We investigate the patterns observed in ejecta curtain induced by hypervelocity impact (2-6 km/s) with a variety of the size and shape of target particles. We characterize the patterns by an angle, defined as the ratio of the characteristic length of the pattern obtained by Fourier transformation to the distance from the impact point. This angle is found to be almost the same as that obtained by the reanalysis of the patterns in the previous study at lower impact velocities (Kadono et al., 2015, Icarus 250, 215-221), which are consistent with lunar crater-ray systems. Assuming that the pattern is formed by mutual collision of particles with fluctuation velocity in excavation flow, we evaluate an angle at which the pattern growth stops and show that this angle is the same in the order of magnitude as the ratio of the fluctuation velocity and the radial velocity. This relation is confirmed in the results of experiments and numerical simulations. Finally, we discuss the dependence of the patterns on impact conditions. The experiments show no dependence of the angle on impact velocity. This indicates that the ratio between the fluctuation and radial velocity components in excavation flow does not depend on impact velocity. Moreover, the independences on particle size and particle shape suggest that the angle characterizing the structure of the patterns does not depend on cohesive force. Since cohesive forces should be related with elastic properties of particles, the structure does not depend on elastic properties, though inelastic collisions are important for the persistence and contrast of the patterns. | astro-ph.EP | astro-ph | Crater-ray formation through mutual collisions of hypervelocity-impact
induced ejecta particles
Toshihiko Kadonoa,*, Ayako I. Suzukib, Rintaro Matsumuraa, Junta Nakaa, Ryo
Suetsugua, Kosuke Kurosawac, and Sunao Hasegawab
a Department of Basic Sciences, University of Occupational and Environmental Health, Kitakyusyu 807-
8555, Japan
b Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, Kanagawa 252-5210,
Japan
c Planetary Exploration Research Center, Chiba Institute of Technology, Chiba 275-0016, Japan
* Corresponding author.
E-mail address: [email protected] (T. Kadono)
1
ABSTRACT
We investigate the patterns observed in ejecta curtain induced by hypervelocity impact
(2−6 km/s) with a variety of the size and shape of target particles. We characterize the
patterns by an angle, defined as the ratio of the characteristic length of the pattern
obtained by Fourier transformation to the distance from the impact point. This angle is
found to be almost the same as that obtained by the reanalysis of the patterns in the
previous study at lower impact velocities (Kadono et al., 2015, Icarus 250, 215-221),
which are consistent with lunar crater-ray systems. Assuming that the pattern is formed
by mutual collision of particles with fluctuation velocity in excavation flow, we evaluate
an angle at which the pattern growth stops and show that this angle is the same in the
order of magnitude as the ratio of the fluctuation velocity and the radial velocity. This
relation is confirmed in the results of experiments and numerical simulations. Finally, we
discuss the dependence of the patterns on impact conditions. The experiments show no
dependence of the angle on impact velocity. This indicates that the ratio between the
fluctuation and radial velocity components in excavation flow does not depend on impact
velocity. Moreover, the independences on particle size and particle shape suggest that the
angle characterizing the structure of the patterns does not depend on cohesive force. Since
cohesive forces should be related with elastic properties of particles, the structure does
not depend on elastic properties, though inelastic collisions are important for the
persistence and contrast of the patterns.
2
1.Introduction
Crater rays are typically observed around craters on solid surfaces of celestial bodies
such
as
the
moon
(see
e.g.,
Kaguya
website:
http://www.kaguya.jaxa.jp/gallery/index_e.html). The nature of the rays has been studied
for a long time, and the origin of the brightness of the rays is fairly understood (e.g.,
Hawke et al., 2004). Another remarkable feature of crater rays, the spatial non-uniformity,
is recently paid attention as a geologic process such that the spatially heterogeneous rays
determine the rate of degradation of small craters (Minton et al., 2019). However, the
mechanism of non-uniformity of crater-rays remains an unsolved problem; there are only
a few studies investigating the mechanism for the generation of non-uniformities of the
rays such as high-velocity detonation products in explosion cratering experiments
analogous to the impact-induced vapor (Andrews, 1977), interaction of shock waves with
old craters (Shuvalov, 2012), pattern formation through mutual collisions in granular
media (Kadono et al., 2015), and, effect of the topography around the impact point
(Sabuwala et al., 2018).
Impact on planetary and lunar surfaces is expected to occur at hypervelocity (> ~km/s).
However, in our previous experiments the results at the impact velocities lower than ~100
m/s were shown (Kadono et al. 2015). The impact velocity was also low in the
experiments by Sabuwala et al. (2018) (< ~10 m/s; projectiles were freely falling).
Therefore, there has been no experimental studies on the non-uniformity of crater-rays at
hypervelocity expected in natural impacts.
In this paper, we investigate the nonuniform structures observed in ejecta curtain
3
induced by hypervelocity impact (2−6 km/s). First, we describe the results of
hypervelocity impact experiments with granular targets with a variety of the particle size
and shape, in particular, the spatial distributions of granular materials during their flight
after impact by taking consecutive images of the ejecta curtain using a high-speed camera.
Then, we analyze the patterns observed in ejecta curtain using Fourier transformation. As
one possible process to form such non-uniform distribution in the ejecta curtain, mutual
collision of particles was suggested (e.g., Lohse et al., 2004; Kadono et al., 2015), based
on the result that inelastic mutual collision of particles plays an important role to
spontaneously form inhomogeneous clustering state (e.g., Luding and Herrmann, 1999;
Goldhirsch, 2003). Therefore, assuming that collision process underlies the pattern
formation, we derive the relation between the patterns and the fluctuation velocity with
the support of numerical simulations and finally discuss the dependence of the patterns
on impact conditions.
2.Experiment
We carried out impact experiments using a two-stage hydrogen-gas gun at Institute of
Space and Astronautical Science, JAXA. Spherical polycarbonate projectiles with a
diameter of 4.8 mm (0.068 g in mass) were accelerated, which impacted perpendicularly
onto the surface of granular targets. Five shots were made: a spherical glass bead target
with a diameter of 0.04−0.05 mm at an impact velocity of 2.3 km/s (denoted as GB50),
spherical glass beads with a diameter of 0.09−0.11 mm at impact velocities of 2.2 km/s
(GB100_2) and 5.9 km/s (GB100_6), spherical glass beads with a diameter of 0.25−0.36
mm at impact velocities of 2.3 km/s (GB300), and irregularly shaped silica (SiO2) sands
4
with a diameter of 50−100 m at an impact velocity of 2.3 km/s (SS70). The targets were
poured into a bowl, which had a radius of 15 cm and depth of 10 cm with a flat bottom,
and set in a vacuum chamber. The ambient pressure was less than 2.5 Pa. The motion of
the ejecta curtain was observed using a high-speed video camera (HPV-X, Shimadzu Co.
Ltd) with a rate of 2 ms per frame. After the shot, we measured the rim-to-rim diameter
of final craters. A schematic diagram of the experimental configuration is shown in Fig.
1.
Fig. 1. Schematic diagram of the experimental configuration. A polycarbonate projectile
impacted perpendicularly onto the target surface from the top in a vacuum chamber. A
high-speed camera and a light were set outside of the chamber.
3.Results
Figure 2 shows consecutive images of GB100_2 (the movie for each shot is available
in the Supplemental Materials (SM)). A projectile coming from the top impacted
5
perpendicularly onto the surface of the granular target and the spatial distribution of the
ejected particles from each shot showed a nonuniform pattern.
Fig. 2. Consecutive images of ejecta curtain for a glass beads target GB100_2. A
polycarbonate projectile impacted perpendicularly onto the target surface from the top.
The upper-left panel shows the ejecta curtain at 34 ms after the impact and the time
proceeds from the upper-left to -right and then lower-left to -right panels with an interval
of 10 ms. A white horizontal line in lower-right panel indicates a spatial scale of 50 mm.
Fig. 3. Intensity distributions at 44 ms after the impact of GB100_6 (black line), GB100_2
(red), GB50 (blue), GB300 (orange) and SS70 (green). The curves are offset from each
other to avoid crossovers.
6
We measured the distribution of intensities along a horizontal line on the ejecta curtain
(the measured lines are shown in Figs. S1a−e in the SM) with a length of 128 pixels (163
mm). Note that the length between two points along the horizontal arcs in the ejecta
curtain is different from that projected on a plane (as images taken by the camera). Hence,
we corrected the intensity data taken by the camera in order to obtain intensities at points
evenly distributed along the horizontal arcs in the ejecta curtain (see the SM). The
corrected distributions are shown in Fig. 3, where the horizontal axis denotes the spatial
length along the horizontal arc in the ejecta curtain. The distributions have an oscillatory
behavior; hence, we analyzed the distributions using Fourier transformation to investigate
their periodic structures. The spectra were plotted for every 10 ms starting from 30 ms
after the impact; each spectrum is an average of the spectra for five consecutive frames
(e.g., 30, 32, 34, 36, and 38 ms) (the spectra for all of the shots are shown in Fig. S3 in
the SM). One example of the obtained power spectra is shown in Fig. 4 for GB100_2,
where the vertical axis denotes the square of the amplitude F and the horizontal axis
denotes the frequency k in units of mm-1. Note that only the data for k between 0.05 and
0.3 mm-1 were plotted. The spectra at large wavenumbers of 49−64 (k > ~0.30 mm-1) were
considered as noise and the counts averaged over this range were subtracted from the data.
In addition, the spectra at small wavenumbers of 1−9 (k < ~0.05 mm-1), corresponding to
~1/10 of the length of the line that we measured, were not considered because they may
include the influence of the length of the line that we measured. The power (F2) in the
spectra decreases with k and does not show clear peaks even though the scatter is large,
which is consistent with the findings of a previous study (Kadono et al., 2015).
7
Fig. 4. A power spectrum for GB100_2 as a function of frequency k (mm-1). This
spectrum represents an average of the spectra for five consecutive frames during 40−48
ms (i.e., 40, 42, 44, 46, and 48 ms) after the impact. The characteristic length c is
obtained by fitting a Gaussian function to the spectra. The fitting curve is also shown in
the figure as a dotted curve.
In the investigation of cluster formation process in a freely evolving gas and the
density correlation function, it is found that the system evolution is described by a single
length scale and that the asymptotic dependence of this length on time is consistent with
diffusive growth (Das and Puri, 2003). Based on this result, we suppose that diffusion is
fundamental for the pattern formation in the ejecta curtain and fit a Gaussian function
AExp[−Bk2] to the power spectra in Fig. 4, where A and B are the fitting parameters.
Parameter B corresponds to c
2, where c is the characteristic length of the patterns of
ejecta curtain. Figure 4 shows the fitting curve alongside the data, which demonstrates
the suitability of the Gaussian function for the experimental data. We consider an "angle"
c at which the impact point takes a view of c, defined as c/r, where r is the distance
8
from the impact point to a point on the horizontal arc that we measured on the ejecta
curtain. Using an average value of r during an interval of 10 ms, c is obtained. Figure 5
shows c as a function of time normalized by the characteristic crater formation time τ =
(Dc/g)1/2, where Dc is the rim-to-rim diameter of the final crater (~13.5−26.0 cm in our
experiments) and g is the acceleration due to gravity (Melosh, 1989). It is noted that τ
becomes ~100 ms in our case, and hence, the results shown correspond to the
intermediary and later stages of crater formation. We also show the result of a previous
study using 100-μm glass beads at a lower impact velocity of ~100 m/s (denoted as
GB100_100), obtained by the reanalysis of the images during 30−90 ms after the impact
shown in Fig. 3 of Kadono et al. (2015) with the same procedure as that in this paper (i.e.,
fitting a Gaussian function to the noise-subtracted power spectra); we exclude the data
from the images after 90 ms because they are slightly out of the focus of the imaging
optics. Results show that c remains almost constant ~0.06 rad across time after the
impact (though the result with GB300 largely scatters at the later stages (> ~0.5) due to
the sparseness of ejecta distribution and the low contrast of images, the average during
the stages is similar to that at the earlier stages). A constant angle implies that the pattern
expands geometrically and does not evolve; the formation process in the ejecta curtain
has finished at the observation time, which is consistent with a previous study (Kadono
et al., 2015). The fact that no difference in c is found among the results of GB100_100,
GB100_2, and GB100_6 indicates that no dependence of c is expected on impact
velocity. Moreover, no difference among the results of GB50, GB100_2, and GB300
suggests that no dependence of c is expected on particle sizes. On the other hand, the
9
results show that c for silica sands appears to be slightly lower than that for glass beads
at the later stages, but the difference of the data seems to be within their scatters. In fact,
the averages of SS70 and GB50 during the common time range shown in Fig. 5 (0.3 < τ
< 0.8) are 0.051±0.009 and 0.065±0.007, respectively, suggesting that there is no
significant difference (the difference in the previous results at lower impact velocities is
also small within the errors (Kadono et al., 2015)).
Fig. 5. Angle θc (in units of radian), defined as the ratio of the characteristic length c to
the distance from the impact point, is plotted as a function of time after impact normalized
by the characteristic crater formation time τ = (Dc/g)1/2. Since c is obtained from the
averaged power spectra for an interval of 10 ms, θc is plotted at the median of each range.
The errors in θc are mainly from those in c caused by the fitting.
4.Discussion
4.1. θc and inelastic collision process
Here, we consider target granular materials in excavation flow and look at a
10
hypothetical horizontal circle getting on the excavation flow centered at impact point with
a radius of r(t), which increases with time t with a velocity of vr(t). When the particles on
the circle having a fluctuation velocity perpendicular to the radial direction (i.e., along
the circle) v(t), cluster formation occurs through inelastic collisions on the circle. Length
scale between clusters on the circle λ(t) increases with time as clustering proceeds. The
experimental result that θc is constant in time implies that cluster growth stops at a finite
time. To represent this situation, we suppose the circumference length of the circle
geometrically becomes so large in time that particle collisions along the circle cease. The
circumference length of the circle L(t) = 2πr(t) increases with time due to the geometrical
expansion as dL(t)/dt = 2πvr(t) and the relative velocity vrel(t) between two points
separated by λ(t) along the circle is vrel(t) = {λ(t)/L(t)}(2πvr(t)) = {λ(t)/r(t)}vr(t).
Supposing that collision due to diffusion stops when v(t) ~vrel(t), we obtain the
characteristic angle θc, at which particle accumulation through collisions stops, as θc =
λ(t*)/r(t*) ~v(t*)/vr(t*), where t* is the time at which the collision stops. This indicates
that θc is the same order of magnitude as the ratio of the fluctuation velocity to the radial
velocity of excavation flow.
4.2. θc versus the ratio of fluctuation velocity to radial velocity
4.2.1. Experiment
In GB300, spatially scattered particles were observed to be apart about an order of δ
~several mm from the ejecta curtain at r ~10 -- 20 cm in distance from the impact point
(Fig. 6). Moreover, in experiments using slightly larger glass spheres (~500 μm), spatially
11
scattered particles were also observed to be δ ~a few mm and r ~10 -- 20 cm (Tsujido et al.,
2015). This scattering was caused by velocity fluctuation in the direction of the zenith
angle vφ. The ratio of the fluctuation velocity vφ to radial velocity vr of particles is
evaluated to be δ/r ~0.01 -- 0.1, which is consistent with a value of ~0.01 -- 0.1 obtained
from the previous experiments with freely falling granular streams (Amarouchene et al.,
2008; Royer et al., 2009). Assuming vφ and vr to be v(t*) and vr(t*), respectively, the
ratio v(t*)/vr(t*) is ~0.01 -- 0.1 and actually the same in the order of magnitude as c
obtained in our experiments as shown in Fig. 5.
Fig. 6. Ejecta curtain in the target of GB300 at 62 ms after the impact. Some scattered
particles can be recognized (indicated by a bold arrow). Horizontal bar indicates 30 mm
and a narrow line shows the distance from the impact point to the ejecta curtain.
4.2.2. Numerical simulation
To confirm alternatively the relation between θc and v(t*)/vr(t*), we numerically
investigated the pattern in impact-induced ejecta curtain using an open-source discrete-
element-method simulator, LIGGGHTS (Kloss et al., 2012), where the particles
considered were soft spheres and the interactions between the particles in contact were
12
taken into account. The normal repulsive forces were represented by a spring and dashpot
in parallel. When two particles collided, the normal component of the velocity, vn,
between two particles was reduced to -- evn, where e was the coefficient of restitution set
to a value of 0.1. We also considered the effect of friction on the change in the tangential
component of the velocity between two particles based on Coulomb's friction law (the
coefficient of friction is set to 0.5) and the gravitational force (the gravitational
acceleration g was set to 9.81 m/s2), but cohesive force was not included in the simulation
for simplicity.
Fig. 7. The results of the numerical simulation. (a) Particle distributions as a function of
azimuthal angle in the ejecta curtain with height between 8.3−12 cm at 40, 60, and 80 ms
after impact. The profiles at 40 and 60 ms are offset to avoid crossover adding +30 and
+10, respectively. (b) Power spectrum of the distribution at 60 ms. Fitting curve is also
shown as a dotted curve. Spectra at 40 and 80 ms are shown in Fig. S4 in the SM. (c)
Characteristic angle θc for the particle distributions as a function of time. The ratio
<va>/<vr> is also plotted. The errors in θc and <va>/<vr> are from those in the fitting and
in the averages over all particles, respectively.
13
The simulation considered 1.14×106 spherical particles with a radius of 0.7 mm as a
target, whose density, Young modulus, and Poisson ratio were set to values of 2.7 g/cm3,
9.4 MPa, and 0.17, respectively. Initially, these particles were accumulated at random in
a hypothetical rectangular box with a size of 20 cm × 20 cm along the horizontal plane
and a height of ~15 cm, and then gravitationally settled in a rectangular box with a size
of 20 cm × 20 cm along the horizontal plane and a height of 7 cm. We considered a
hypothetical sphere with a radius of 1.3 cm as projectile, in which 103 spherical particles
with a radius of 0.7 mm (whose density, Young modulus, and Poisson ratio were set to
values of 1.0 g/cm3, 9.4 MPa, and 0.17, respectively) were accumulated. This sphere
collided vertically at the center of the target surface with a velocity of 100 m/s (see the
SM for an animation). We considered the distribution of particles in the ejecta curtain
between the heights of 8.3−12 cm from the bottom surface of the target box. The ejecta
curtain was divided azimuthally into 512 boxes. A part of the distributions at 40, 60, and
80 ms after the impact is shown as a function of the azimuthal angle (Fig. 7a). We
performed Fourier transformation, and the power spectrum was obtained as shown in Fig.
7b as a function of frequency k in units of rad-1 (the spectra at large k > 12.8 rad-1
(wavenumbers > 80) and small k < 1.5 rad-1 (wavenumbers of 1−9) are not considered)
(Each power spectrum of the distributions shown in Fig. 7a at 40, 60 and 80 ms is shown
in Fig. S4 in the SM). Though the spectrum fairly scatters, the curve appears to decrease
with k. The characteristic angle θc was obtained by fitting a Gaussian function (a fitting
curve is shown in Fig. 7b) and compared with the ratio <va>/<vr>, where <va> and <vr>
14
are the absolute azimuthal and radial velocity components averaged over all particles,
respectively. We considered the average of the absolute azimuthal component <va> as the
fluctuation velocity v. Figure 7c shows θc and <va>/<vr> as a function of time
normalized by a characteristic crater formation time τ = (Dc/g)1/2, where crater (rim-to-
rim) diameter Dc was ~15.5 cm in this simulation. It appears that θc is almost constant
with time at intermediary stages of crater formation; this is consistent with the
experimental result. The ratio <va>/<vr> is also almost constant and ~1/3−1/4 of θc. This
confirms that θc and the ratio between the fluctuation and radial velocity components are
in the same order of magnitude.
4.3. Dependence of θc on impact conditions
The experimental result shows no dependences of the angle c on impact velocity. In
cratering, impact causes a shock wave and, after the initial shock wave has dissipated, the
excavation flow field is established (e.g., Melosh, 1985; 1989; Wada et al., 2006). Though
the excavation flow field created by the shock wave's passage should be complex, our
experimental result indicates that the ratio between the fluctuation and radial velocity
components in the excavation flow field is quite similar regardless of impact velocity.
No difference in c among the results of GB50, GB100_2, and GB300 reveals no
dependence on particle sizes. Furthermore, the difference of c between SS70 and GB50
is small within the scattering of the data (Fig. 5). Therefore, the effect of particle shape is
also small. Since cohesive forces such as van der Waals and capillary forces between
particles depend on size and shape of particles, these results imply that the cohesive forces
and closely related elastic properties of particles do not influence θc representing the
15
distance between clusters, i.e., the structure of patterns. However, this does not mean that
inelastic collision between particles is not important. If the coefficient of restitution e is
not sufficiently low, the particles diffuse and the pattern changes temporary; i.e., e plays
key roles in the formation of distinct and persistent clusters. In fact, in freely falling
granular streams in a vacuum (Möbius, 2006; Royer et al., 2009), Royer et al. (2009)
show that cohesive forces are responsible for the occurrence of discrete, compact, and not
transient clusters and, in experiments using spherical particles rolling on a smooth surface
driven by a moving wall (Kudrolli et al., 1997), energy dissipation due to inelastic
collisions plays a role in the formation of not dispersed clusters. Thus, we conclude that
the elastic properties of particles determine whether distinct and persistent clusters are
formed or not but have little influence on the structure of consequent patterns.
In this paper we used identical particles in size in each shot but the surface of natural
bodies would not consist of identical particles in size but the mix of various sized particles.
When target granular materials include larger particles, the fluctuation velocity of
particles in the excavation flow would be different and also coalescence of particles
should be promoted due to their larger collision cross-section. As a result, the patterns
may be different. In fact, the recent experiments with the target consisting of two kinds
of particles in size (0.1 mm and 1 or 4 mm) mixed at almost the same weight % show that
large inclusions disturb excavation flow as obstacles and cause different flow patterns
such as drags and spurts, but these patterns are temporary and not periodic (Kadono et al.
2019). On the other hand, when the amount of small particles is relatively large, the
pattern would become similar to the one observed in our experiments. However, the
16
systematic investigation of the effects of size distributions is insufficient and hence,
should be done as a future work.
5.Conclusion
We investigated the patterns observed in ejecta curtain induced by hypervelocity
impacts and the dependences of the patterns on the impact conditions such as impact
velocity, particle size, and shape. We considered a quantity that characterizes the pattern,
the angle θc defined as the ratio of the characteristic length of the pattern obtained by
Fourier transformation to the distance from the impact point. This angle θc was found to
be constant through time after the impact, suggesting that the pattern formation finished
by the observation time. We evaluated an angle at which cluster growth stopped assuming
that mutual collision of particles with fluctuation velocity resulted in clusters and showed
that this angle and the ratio of the fluctuation velocity to the radial velocity were in the
same order of magnitude. We confirmed this relation in the results of experiments and the
numerical simulations. The experiments also show no dependence of θc on impact
velocity. This indicates that the ratio between the fluctuation and radial velocity
components in excavation flow does not depend on impact velocity. Moreover, the
independences on particle size and particle shape suggest that the cohesive forces would
not contribute the structure of the patterns.
The formation process of the pattern observed in the ejecta curtain can be summarized
by the following: (1) after the initial shock wave dissipated and the excavation flow field
was established, the particles acquired fluctuation velocity in the horizontal direction and
17
accumulated through inelastic mutual collision, (2) clustering stopped when a
characteristic angle of the pattern became at a certain value, which was the same in the
order of magnitude as the ratio between the fluctuation and radial velocities, and (3) the
pattern expanded geometrically and was observed in the ejecta curtain.
The patterns in ejecta curtain induced by hypervelocity impacts are consistent with
those in the previous study at lower impact velocities (Kadono et al. 2015). Since the
previous results are consistent with lunar crater-rays, we can conclude that the patterns
observed in ejecta curtain induced by hypervelocity impacts are also consistent with the
lunar crater-rays. Crater-rays are also observed on other airless bodies such as Mercury
(e.g., MESSENGER website: http://messenger.jhuapl.edu/Explore/Science-Images-
Database/By-Topic/topic-64.html) and asteroids (e.g., Williams et al. 2014). The
comparison between the ray patterns on such bodies would verify the formation process
that we proposed. For evaluation of such natural crater-ray systems on complex
topography, however, the quantitative methods capable of analyzing the patterns in two-
dimensions would be necessary to assess the similarity and difference between the
patterns of ejecta deposit more precisely than our analysis in one-dimensional.
Acknowledgments
We thank M. Arakawa for the discussion based on his unpublished data and M. Koga
for supporting data analysis and creating some figures. The authors are also grateful to
two anonymous reviewers for helpful comments. This work was supported by
ISAS/JAXA as a collaborative program with the Hypervelocity Impact Facility.
18
19
References
Amarouchene, Y., Boudet, J.-F., Kellay, H., 2008. Capillarylike fluctuations at the interface of falling
granular jets. Phys. Rev. Lett. 100, 218001. http://doi.org/10.1103/PhysRevLett.100.218001.
Andrews, R. J., 1977. Characteristics of debris from small-scale cratering experiments. In: Roddy, D. J.,
Pepin, R. O., and Merrill, R. B. (Eds.), Impact and Explosion Cratering. Pergamon Press, New York,
pp. 1089-1100.
Das, S.K., Puri, S., 2003. Kinetics of inhomogeneous cooling in granular fluids. Phys. Rev. E 68, 011302.
http://doi.org/10.1103/PhysRevE.68.011302.
Goldhirsch, I., 2003. Rapid granular flows. Annu. Rev. Fluid Mech. 35, 267-293.
http://doi.org/10.1146/annurev.fluid.35101101.16114.
Hawke, B. R., Blewett, D. T., Lucey, P. G., Smith, G. A., Bell, J. F., Campbell, B. A., Robinson, M. S.,
2004. The origin of lunar crater rays. Icarus 170, 1-16, 10.1016/j.icarus.2004.02.013.
Kadono, T., Suzuki, A.I., Wada, K., Mitani, N.K., Yamamoto, S., Arakawa, M., Sugita, S., Haruyama, J.,
Nakamura, A.M., 2015. Crater-ray formation by impact-induced ejecta particles. Icarus 250, 215-221.
http://doi.org./10.1016/j.icarus.2014.11.030.
Kadono, T., Suetsugu, R., Arakawa, D., Kasagi, Y., Nagayama, S., Suzuki, A.I., Hasegawa, S., 2019.
Pattern of Impact-Induced Ejecta from Granular Targets with Large Inclusions. Astrophys. J. Lett. 880,
L30. http://doi.org./10.3847/2041-8213/ab303f.
Kloss, C., Goniva, C., Hager, A., Amberger, S., Pirker, S., 2012. Models, algorithms and validation for
opensource DEM and CFD-DEM. Prog. Comput. Fluid Dy. 12, 140-152.
http://doi.org./10.1504/PCFD.2012.047457.
Kudrolli, A., Wolpert, M., Gollub, J.P., 1997. Cluster formation due to collisions in granular material.
20
Phys. Rev. Lett. 78, 1383-1386.
Lohse, D., Bergmann, R., Mikkelsen, R., Zeilstra, C., van der Meer, D., Versluis, M., van der Weele, K.,
van der Hoef, M., Kuipers, H., 2004. Impact on soft sand: Void collapse and jet formation. Phys. Rev.
Lett. 93, 198003. http://doi.org/10.1103/PhysRevLett.93.198003.
Luding, S., Herrmann, H.J., 1999. Cluster-growth in freely cooling granular media. Chaos 9, 673-681.
http://doi.org/10.1063/1.166441.
Melosh, H. J., 1985. Impact cratering mechanics: Relationship between the shock wave and excavation
flow. Icarus 62, 339-343.
Melosh, H. J., 1989. Impact ratering: A geologic process, Oxford University Press, New York, New
York.
Minton, D.A., Fassett, C.I., Hirabayashi, M., Howl, B.A., Richardson, J.E., 2019. The equilibrium size-
frequency distribution of small craters reveals the effects of distal ejecta on lunar landscape
morphology. Icarus 326, 63-87. http://doi.org/10.1016/j.icarus.2019.02.021.
Möbius, M.E., 2006. Clustering instability in a freely falling granular jet. Phys. Rev. E 74, 051304.
http://doi.org./10.1103/PhysRrevE.74.051304.
Royer, J.R., Evans, D.J., Oyarte, L., Guo, Q., Kapit, E., Möbius, M.E., Waitukaitis, S.R., Jaeger, H.M.,
2009. High-speed tracking of rupture and clustering in freely falling granular streams. Nature 495,
1110-1113. http://doi.org./10.1038/nature081155.
Sabuwala, T., Butcher, C., Gioia, G., Chakraborty, P., 2018. Ray systems in granular cratering. Phys.
Rev. Lett. 120, 264501. http://doi.org./10.1103/PhysRrevLett.120.264501.
Shuvalov, V., 2012. A mechanism for the production of crater rays. Meteo. Planet. Sci. 47, 262-267,
10.1111/j.1945-5100.2011.01324.x.
21
Tsujido, S., Arakawa, M., Suzuki, A.I., Yasui M., 2015. Ejecta velocity distribution of impact craters
formed on quartz sand:Effect of projectile density on crater scaling law. Icarus 262, 79-92.
http://doi.org./10.1016/j.icarus.2015.08.035.
Wada, K., Senshu, S., Matsui, T., 2006. Numerical simulation of impact cratering on granular material.
Icarus 180, 528-545. http://doi.org./10.1016/j.icarus.2005.10.002.
Williams, D. A., Denevi, B. W., Mittlefehldt, D. W., et al., 2014. The geology of the Marcia quadrangle of
asteroid Vesta: Assessing the effects of large, young craters. Icarus 244, 74-88.
http://doi.org./10.1016/j.icarus.2014.01.033.
22
|
1301.5101 | 1 | 1301 | 2013-01-22T08:25:45 | The Resolved Asteroid Program - Size, shape, and pole of (52) Europa | [
"astro-ph.EP"
] | With the adaptive optics (AO) system on the 10 m Keck-II telescope, we acquired a high quality set of 84 images at 14 epochs of asteroid (52) Europa on 2005 January 20. The epochs covered its rotation period and, by following its changing shape and orientation on the plane of sky, we obtained its triaxial ellipsoid dimensions and spin pole location. An independent determination from images at three epochs obtained in 2007 is in good agreement with these results. By combining these two data sets, along with a single epoch data set obtained in 2003, we have derived a global fit for (52) Europa of diameters (379x330x249) +/- (16x8x10) km, yielding a volume-equivalent spherical-diameter of 315 +/- 7 km, and a rotational pole within 7 deg of [RA; Dec] = [257,+12] in an Equatorial J2000 reference frame (ECJ2000: 255,+35). Using the average of all mass determinations available forEuropa, we derive a density of 1.5 +/- 0.4, typical of C-type asteroids. Comparing our images with the shape model of Michalowski et al. (A&A 416, 2004), derived from optical lightcurves, illustrates excellent agreement, although several edge features visible in the images are not rendered by the model. We therefore derived a complete 3-D description of Europa's shape using the KOALA algorithm by combining our imaging epochs with 4 stellar occultations and 49 lightcurves. We use this 3-D shape model to assess these departures from ellipsoidal shape. Flat facets (possible giant craters) appear to be less distinct on (52) Europa than on other C-types that have been imaged in detail. We show that fewer giant craters, or smaller craters, is consistent with its expected impact history. Overall, asteroid (52) Europa is still well modeled as a smooth triaxial ellipsoid with dimensions constrained by observations obtained over several apparitions. | astro-ph.EP | astro-ph | The Resolved Asteroid Program - Size, shape, and pole of (52) Europa✩
W. J. Merlinea , J. D. Drummondb, B. Carryc,d, A. Conrade,f , P. M. Tamblyna , C. Dumasg , M. Kaasalainenh , A. Eriksoni,
S. Mottolai , J. Durechj , G. Rousseauk , R. Behrendk,l, G. B. Casalnuovok,m, B. Chinagliak,m , J. C. Christoun , C. R. Chapmana ,
C. Neymanf
a Southwest Research Institute, 1050 Walnut St. # 300, Boulder, CO 80302, USA
b Star fire Optical Range, Directed Energy Directorate, Air Fo rce Research Laboratory, Kirtland AFB, NM 87117-577, USA
c IMCCE, Observatoire de Paris, CNRS, 77 av. Denfert Rochereau, 75014 Paris, France
d European Space Astronomy Centre, ESA, P.O. Box 78, 28691 Villanueva de la Ca nada, Madrid, Spain
eMax-Planck-Institut f ur Astronomie, K onigstuhl, 17, Heidelberg, Germany
fW.M. Keck Observatory, 65-1120 Mamalahoa Highway, Kamuela, HI 96743, USA
g ESO, Alonso de C ´ordova 3107, Vitacura, Casilla 19001, Sant iago de Chile, Chile
h Tampere University of Technology, P.O. Box 553, 33101 Tampere, Finland
i Institute of Planetary Research, DLR, Rutherfordstrasse 2, 12489, Berlin, Germany
j Astronomical Institute, Faculty of Mathematics and Physics, Charles University in Prague, V Hole sovi ck ´ach 2, 1800 0 Prague, Czech Republic
kCdR & CdL group: Lightcurves of minor planets and variable stars
lGeneva Observatory, 1290 Sauverny, Switzerland
mEurac Observatory, Bolzano
nGemini Observatory, 670 N. Aohoku Place, Hilo, Hawaii, 96720, USA
Abstract
With the adaptive optics (AO) system on the 10 m Keck-II telescope, we acquired a high quality set of 84 images at 14 epochs
of asteroid (52) Europa on 2005 January 20, when it was near opposition. The epochs covered its 5.63 h rotation period and, by
following its changing shape and orientation on the plane of sky, we obtained its triaxial ellipsoid dimensions and spin pole location.
An independent determination from images at three epochs obtained in 2007 is in good agreement with these results. By combining
these two data sets, along with a single epoch data set obtained in 2003, we have derived a global fit for (52) Europa of diame ters
a × b × c = (379 × 330 × 249) ± (16 × 8 × 10) km, yielding a volume-equivalent spherical-diameter of 3√abc = 315 ± 7 km, and
a prograde rotational pole within 7◦ of [RA; Dec] = [257◦; +12◦ ] in an Equatorial J2000 reference frame (Ecliptic: 255◦; +35◦ ).
Using the average of all mass determinations available for (52) Europa, we derive a density of 1.5 ± 0.4 g cm−3 , typical of C-type
asteroids. Comparing our images with the shape model of Michalowski et al. (Astron. and Astrophys. 416, p 353, 2004), derived
from optical lightcurves, illustrates excellent agreement, although several edge features visible in the images are not rendered by
the model. We therefore derived a complete 3-D description of (52) Europa’s shape using the KOALA algorithm by combining
our 18 AO imaging epochs with 4 stellar occultations and 49 lightcurves. We use this 3-D shape model to assess these departures
from ellipsoidal shape. Flat facets (possible giant craters) appear to be less distinct on (52) Europa than on other C-types that
have been imaged in detail, (253) Mathilde and (511) Davida. We show that fewer giant craters, or smaller craters,
is consistent
with its expected impact history. Overall, asteroid (52) Europa is still well modeled as a smooth triaxial ellipsoid with dimensions
constrained by observations obtained over several apparitions.
3
1
0
2
n
a
J
2
2
]
P
E
.
h
p
-
o
r
t
s
a
[
1
v
1
0
1
5
.
1
0
3
1
:
v
i
X
r
1
a
Keywords:
1. Introduction
2
3
4
5
6
Direct, accurate measurements of asteroid shapes, sizes, and
pole positions are now possible for larger asteroids that can be
spatially resolved using the Hubble Space Telescope (HST)
or
large ground-based telescopes equipped with adaptive
optics (AO). Physical and statistical study of asteroids requires
✩Based on observations at the W. M. Keck Observatory, which is operated as
a scientific partnership among the California Institute of T echnology, the Uni-
versity of California and the National Aeronautics and Space Administration.
The Observatory was made possible by the generous financial s upport of the W.
M. Keck Foundation.
Email address: [email protected] (W. J. Merline)
7
accurate knowledge of these parameters.
Improved sizes
permit improved estimates of albedo, in turn allowing better
interpretation of surface composition. In those cases where we
have an estimate of the mass, for example from the presence
of a satellite, the uncertainty in the volume of the asteroid is
the overwhelming uncertainty in attempts to derive its density
(Merline et al. 2002). Of course, density is the single most
critical observable having a bearing on bulk composition,
porosity, and internal structure (Merline et al. 2002; Britt et al.
2002, 2006). With our technique of determining the size of an
asteroid by following its changing apparent size, shape, and
orientation, the uncertainties in volume can now be reduced
to the level of the mass uncertainty, vastly improving our
8
9
10
11
12
13
14
15
16
17
18
19
Preprint submitted to Icarus
December 12, 2013
Figure 1: Apparent angular sizes of Solar System objects. Asteroid, moon, comet, and TNO diameters are plotted against their geocentric distances, defined as the
difference between their semi-major axis and 1 AU. Symbol size corresponds to physical diameter. Gray scales represent the changing apparent size with geocentric
distance. A body of a given size moves along the oblique lines as its distance from the Earth changes. The angular resolutions at CFHT, Keck and future TMT and
E-ELT are also shown for different filters (V: 0.6 µm, and K: 2.2 µm). Typical NEA populations (Apollos, Atens, and Amors) are also shown, as represented by
(1566) Icarus, (99942) Apophis, and (433) Eros, respectively.
34
con fidence in the derived asteroid densities. The improveme nt
comes about because we can see the detailed shape, track edge
or surface features during rotation, and often can make an
immediate pole determination.
Dedicated study of asteroids now allows directly observable
shape pro files, and already has shown that some asteroids sho w
large departures from a reference ellipsoid that may provide
clues to the body’s response to large impacts over time (e.g.,
(4) Vesta, Thomas et al. 1997). For asteroid (511) Davida, we
suggested (Conrad et al. 2007) that such features (e.g., large
flat facets) may be analogs of the giant craters, seen edge-on ,
in the images of (253) Mathilde during the NEAR mission
(Veverka et al. 1999) flyby. If giant craters are evident on these
surfaces, they can be related to the impact history and impact
flux over time, and there is some chance they can be associated
with asteroid families or clusters that are being identified by
numerical back-integration and clustering of orbital elements
(e.g., Nesvorn ´y et al. 2002 ).
As we have demonstrated with asteroid (511) Davida
(Conrad et al. 2007), we can derive an asteroid’s triaxial ellip-
soid dimensions and rotational pole location in a single night.
However, we now have developed the ability to combine sets
of similar observations obtained at different viewing aspects
to make a global fit to all of the images, drastically reducing
dimension uncertainties that might be due to sparse rotational
sampling or peculiar observing geometries (Drummond et al.,
in preparation). The leverage of widely spaced observations
and the accompanying range of viewing aspects allows un-
precedented accuracy in derived parameters. We can then use
these estimates to project the apparent size and shape of an
asteroid into the past or future, making the asteroid useful as a
reference or calibration object.
Here we report on the physical properties of the asteroid
(52) Europa as a part of our on-going Resolved Asteroid Pro-
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
2
45
37
38
39
42
40
41
35
36
46
47
43
44
gram. We routinely image the apparent disk of asteroids, and
20
search their close vicinity for companions, aiming at setting
better constraints on their spin-vector coordinates, 3-D shapes,
sizes, and multiplicity. One of our main goals is to derive
(or better constrain) their densities. We use two independent
methods to determine size, shape, and pole position of the
target asteroids. One of these is based on the assumption
that the shape is well-described by a smooth triaxial ellipsoid
(see Drummond 2000; Drummond et al. 2009a, 2010,
for
instance). Our other method allows construction of full 3-D
shape models by combining our AO images with other data
types, when available (e.g., optical lightcurves and stellar
occultations, see Carry et al. 2010a,b), in the technique we call
KOALA (Knitted Occultation, Adaptive-optics and Lightcurve
Analysis, see Carry et al. 2010a; Kaasalainen 2011).
48
resolution, approximated by θ = λ/D 49
The best angular
(radian), with λ the wavelength and D the diameter of the
50
telescope aperture, of current ground-based optical telescopes
is about 0.04′′ (Keck/NIR). Due to systematics, however, we
have found that our ability to accurately measure sizes and
details of the apparent shape degrades below about 0.10′′,
based on simulations and observations of the moons of Saturn
and simulations (Carry 2009; Drummond et al. 2009b). The
sample of observable asteroids (i.e., having angular sizes that
get above about 0.10′′ ) is therefore limited to about 200.
This limit
in angular resolution can be converted to a
physical diameter. As can be seen in Fig. 1, we can probe the
size distribution of main-belt asteroids down to about 100 km,
while Pluto is the only Trans-Neptunian Object (TNO) whose
apparent disk can be resolved. At opportune times, we have
been able to resolve the disks of Near-Earth Asteroids (NEAs,
for example, see Merline et al. 2011, 2012). The next genera-
tion of optical facilities will allow an improvement in angular
66
resolution by a factor of 3-4 due to mirror size alone (30 m 67
64
65
54
55
62
63
56
57
51
52
59
60
53
58
61
for TMT and 40 m for E-ELT), allowing the observation of
more than 500 asteroids, even if we consider only objects that
reach half (or 0.05′′ ) of the current size limits (We computed
the expected apparent diameter of asteroids for the 2020 –20 30
period, and counted objects when apparent diameters reach
0.05′′ within this period.)
Second-generation instruments
with high-Strehl AO corrections into visible wavelengths are
planned for these large ground-based telescopes, providing
another factor of 5 improvement due to operation at shorter
wavelengths.
these two factors should provide
Together,
more than an order-of-magnitude improvement with respect
to current resolution. Almost 7 000 asteroids should then
be observable with apparent diameters greater than 0.01′′ .
This breakthrough in imaging capabilities will also enable
the spatial resolution of apparent disks of TNOs larger than
500 km, larger moons (∼100 km) of Uranus and Neptune, small
moons of Jupiter and Saturn, main-belt asteroids of few tens of
kilometers, and NEAs of several hundred meters in favorable
conditions (Fig. 1).
2. Disk-resolved imaging observations
For asteroid (52) Europa, our primary data set was taken on
2005 January 20. In addition, we observed (52) Europa at one
epoch on 2003 October 12, and at three epochs on 2007 May 28.
In 2005 we obtained adaptive optics images of (52) Europa at
H (1.6 µm) and Kp (2.1 µm) bands with NIRC2 (van Dam et al.
2004) on the Keck II 10 m telescope, and give the observing log
in Table 1. The 2003 and 2005 images were taken using the first
generation Keck wave-front controller; the 2007 images were
taken using Keck’s next generation wavefront controller (NG-
WFC, van Dam et al. 2007) under similar conditions. Strehl ra-
tios were 30%, 27%, and 40% on average, respectively, for the
2003, 2005, and 2007 epochs. The latter, higher, value re flec ts
the NGWFC changes which, in addition to a new detector, in-
clude improvements to the electronics and to the software. The
data set consists of 111 images: 9 from 2003, 84 from 2005,
and 18 from 2007, that result in 18 composite images (Table 1).
Although less extensive and at a larger distance from Earth,
the 2007 data add an important new epoch to our 2005 data. By
combining all 3 data sets (2003, 2005, and 2007), our goal was
to derive a global fit that spans a wide range of viewing geome-
tries and provide tight constraints on the size, shape, and pole
for (52) Europa.
When observing at Kp in good seeing conditions, adaptive
optics on Keck II delivers diffraction-limited resolution ele-
ments of width approximately 50 milli-arcsecond (mas). We
used the narrow plate-scale (9.942 ± 0.050 mas/pixel) of the
NIRC2 camera, oriented North-up (±0.15◦ , Konopacky et al.
2007) for all the observations.
3. Triaxial Ellipsoid (TE) Assumption
3.1. 2005 January 20
Each of seven sets of six H-band images and seven sets of six
Kp-band images of asteroid (52) Europa obtained in 2005 was
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
3
53
63
66
57
58
59
60
55
56
54
sky-subtracted, and then fit in the Fourier plane for the aste r-
68
oid and Lorentzian PSF, using our method of Parametric Blind
Deconvolution (PBD, as described by Drummond et al. 1998;
Drummond 2000; Conrad et al. 2007). Asteroid ellipse param-
eters were computed as weighted means from each set of six im-
ages obtained at each filter and each rotational phase or epoc h.
These ellipse parameters (apparent major axis length α, minor
axis length β, and an orientation angle PAα ), were then used to
convert the series of apparent diameters and orientations to the
full triaxial-ellipsoid diameters and direction of (52) Europa’s
rotational pole through a non-linear least squares inversion (see
Drummond 2000, for instance). The results of the fit are given
in Table 2.
In addition to the direct PBD methodology, as cross-checks,
we use two additional avenues to get to the triaxial-ellipsoid
solutions.
In the first of these, the data were flat- fielded,
shifted, and added at each rotational epoch (Fig. 2), and a sin-
69
gle deconvolved image was created with the Mistral algorithm 70
(Mugnier et al. 2004), for each epoch and each filter. These
71
seven Kp and seven H deconvolved images (Fig. 3) were again
fit in the Fourier plane for their apparent ellipse parameter s, and
the series was fit for the full triaxial solution, also given i n Ta-
ble 2.
64
65
72
73
61
62
67
68
74
75
Figure 2:
images of
shifted, and added,
flat-fielded,
Sky-subtracted,
(52) Europa, from 2005, before deconvolution, rotated so that the asteroid’s
spin axis is vertical. Although the direction to the Sun is indicated, the solar
phase angle was only 5.5◦ , making the Sun nearly perpendicular to the plane
of the figure. The rotational phase in degrees, ± 360◦ , of each tile is placed on
top of it for placement in Fig 5. The Kp-band images are in the first and third
columns while the H-band images always follow by a few degrees rotation in
the second and fourth columns.
76
Finally, ellipse parameters were derived from fitting the
77
edges produced by a Laplacian of Gaussian wavelet transform 78
(Carry et al. 2008) on the Mistral deconvolved images. A full
79
triaxial solution can then be found from these ellipse parame-
ters, and is given in Table 2.
The adopted triaxial solution
for (52) Europa, independently determined from the 2005 data,
81
82
80
Date
(UT)
2003-10-12 - 11:48
2005-01-20 - 10:39
2005-01-20 - 10:43
2005-01-20 - 11:25
2005-01-20 - 11:28
2005-01-20 - 12:02
2005-01-20 - 12:04
2005-01-20 - 13:01
2005-01-20 - 13:04
2005-01-20 - 13:45
2005-01-20 - 13:48
2005-01-20 - 14:16
2005-01-20 - 14:18
2005-01-20 - 15:02
2005-01-20 - 15:05
2007-05-28 - 11:44
2007-05-28 - 12:54
2007-05-28 - 13:01
∆
(AU)
3.02
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
2.79
3.41
3.41
3.41
r
(AU)
2.07
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
1.84
2.69
2.69
2.69
φ
(◦ )
7.2
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
5.5
13.3
13.3
13.3
mV
(mag)
10.8
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
10.3
11.9
11.9
11.9
ϕ
(′′ )
0.25
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.28
0.19
0.19
0.19
Rotation phase
(◦ )
26
6
9
55
58
95
97
157
160
204
206
237
239
-74
-71
105
179
186
Filter
Kp
Kp
H
Kp
H
Kp
H
Kp
H
Kp
H
Kp
H
Kp
H
Kp
Kp
Kp
Table 1: Observing log: heliocentric distance (∆), range to observer (r), phase angle (φ), visual apparent magnitude (mV ), angular diameter (ϕ), and arbitrary
rotation phase (zero phase being defined for a lightcurve max imum, i.e., when the apparent cross-section of (52) Europa is the largest) for each epoch (reported in
UT).
Parameter
a (km)
b (km)
c (km)
SEPβ (◦ )
PAnode (◦ )
ψ0 (UT)
EQJ2000 (α0 ,δ0 in ◦ )
σ radius (◦ )
ECJ2000 (λ0 ,β0 in ◦ )
PBD
377 ± 3
331 ± 3
236 ± 9
+27 ± 3
339 ± 1
10.35 ± 0.03
261;+10
1
260;+34
Mistral
376 ± 3
332 ± 3
246 ± 8
+25 ± 3
339 ± 1
10.33 ± 0.03
260;+11
1
258;+34
Edges
381 ± 4
335 ± 4
249 ± 10
+25 ± 5
338 ± 1
10.28 ± 0.04
259;+12
1
257;+35
Mean
378 ± 3
332 ± 3
244 ± 8
+25 ± 3
338 ± 1
10.30 ± 0.03
260;+11
1
258;+34
Table 2: Triaxial-ellipsoid parameters for our 2005 data, with three different data-processing methods: PBD images, Mistral deconvolved images, and edge fitting.
The average values for the parameters are reported in the last column. The quantities derived from the fits of the 2005 data are: triaxial ellipsoid diameters a, b, and
c; the sub-Earth latitude SEPβ ; the line of nodes (the intersection of the asteroid’s equator and the plane of the sky) PAnode ; and the UT of the instant when the
long axis a lies in the plane of the sky along the line of nodes ψ0 . Uncertainties reported here are formal error bars of the fit , see the text for a discussion on the
systematics.
1
2
3
4
5
6
7
8
9
10
11
12
13
is derived from the series of mean ellipse parameters at each
epoch, that is, from the mean of the PBD images, the Mistral
deconvolved images, and the edge- fitting at each epoch. This
preferred mean fit is plotted against observations in Fig. 5. The
location of the pole on the Ecliptic globe is shown in Fig. 6,
along with the locations derived from lightcurves analysis by
others.
Our imaging of (511) Davida (Conrad et al. 2007) showed
large edge features that could be followed during rotation, even
in the raw images. While there may be similar features on
(52) Europa, they do not appear as consistently in the edge pro-
files and are not as easy to track. The features are not as large
or prominent as those on Davida, relative to our reference ellip-
soid. Later in the paper, we use 3-D shape modeling to try to
study these departures from a pure ellipsoid shape.
3.2. 2007 May 28
We also acquired AO observations of (52) Europa at Keck
in 2007 (Table 1). Following the recipe from the last section,
we formed the mean apparent parameters from the three meth-
ods already described (PBD, deconvolved images, and outlines
from the deconvolved images). Although not expected to yield
significant results because the three 2007 observations pro vide
only nine observables to find six unknowns, we nevertheless
fit the three observations for a triaxial ellipsoid (Table 3 and
Fig. 7), and found that the model is in surprisingly good agree-
ment with the results from the 2005 set in Table 2.
14
15
16
17
18
19
20
21
22
23
24
25
26
4
9
400
10
20−Jan−2005 UT
12
11
13
14
0.3
0.25
0.2
c
e
S
c
r
A
−45
0
45
90
Rotational Phase
135
180
225
270
Conv
Deconv
Edges
)
m
k
(
r
e
t
e
m
a
i
D
350
300
250
200
−90
20
s
i
x
A
g
n
o
L
f
o
A
P
350
320
290
−90
−45
0
45
90
Rotational Phase
135
180
225
270
Figure 5: Triaxial ellipsoid fit to measured ellipse paramet ers for our 2005
data. In the upper subplot, each image’s long and short axis dimensions are
plotted along the upper and lower lines, respectively. The H-band epochs follow
the Kp-band epochs by a few degrees, and the different symbols represent the
different methods used to extract the ellipse parameters (PBD or Conv, Mistral
or Deconv, and Edges). The solid lines are the prediction for the projected
(full) ellipses from the mean triaxial ellipsoid parameters (Table 2). The dashed
lines are for the ellipse parameters for the terminator ellipse, which, because
the solar phase angle is only 5.5◦ , fall on the solid lines. The data should lie
approximately midway between the dashed and solid lines (here, that means on
the coincident solid/dashed lines). The lower subplot shows the position angle
of the long axis (PAα ) with the same conventions.
+60
180
225
270
+30
315
0
−30
−60
Figure 6: Pole locations for (52) Europa on the Ecliptic globe. The two cir-
cles denote the uncertainty areas around the pole found for 2005 (larger) and
2007 (smaller), while X’s show the positions found from previous workers us-
ing lightcurves.
-73
-71
Figure 3: Same as in Fig. 2 for the Mistral deconvolved images of (52) Europa.
Rot Phase = −90
Rot Phase = 0
2003
2005
2007
Figure 4: Plane-of-sky orientation of (52) Europa as seen during the 3 observing
dates analyzed. The grids are in equatorial coordinates, with north up, east left.
The blue square is the subsolar point and the red circle is the sub-Earth point.
Two views for each are shown: the maximum (Rot Phase = 0) cross-section
and the minimum (Rot Phase = -90) cross-section for that date. These phases
are the same as those listed in the tables and Figs. 5, 7, and 8. The bold dotted
line represents the line defined as longitude = 0, according to IAU convention
(see Archinal et al. 2011). The longitude is related to the rotational phase by:
longitude = 270◦ - Rot phase. The sense of rotation is given by the right-hand
rule here, with the (positive) pole always northward, and can be discerned in
the figure from the advancement of the bold dotted line by 90 ◦ .
1
2
3
4
5
6
7
3.3. 2003 October 12
The single set of AO images of (52) Europa taken in 2003
(Table 1) does not allow an independent fit for a triaxial solution
because it only provides three observables for six unknowns.
We use these early Keck AO images, however, in a global fit in
the next section. Fig. 8 shows the global fit prediction for the
2003 epoch, together with those data.
3.4. A global solution for all epochs
We can tie the 2003, 2005, and 2007 observations of
(52) Europa together into one simultaneous global fit (Drum-
mond et al., 2012, in preparation), using the sidereal period of
P s = 0.2345816 days (with an uncertainty of 2 in the last digit)
derived by Micha łowski et al. (2004). Along with the global
solution for the triaxial dimensions and pole in Table 4, we list
8
9
10
11
12
13
14
5
9
10
11
12
13
28−May−2007 UT
0.2
c
e
S
c
r
A
0.15
0
45
90
Rotational Phase
135
180
225
270
10
11
12
13
−45
9
Conv
Deconv
Edges
400
350
300
250
)
m
k
(
r
e
t
e
m
a
i
D
200
−90
260
230
200
s
i
x
A
g
n
o
L
f
o
)
°
(
A
P
170
−90
−45
0
45
90
Rotational Phase
135
180
225
270
Figure 7: Same as Fig. 5, but for 2007. The maximum that occurs at 9.74± 0.01,
lighttime corrected, is the same hemisphere as the maximum that occurs at
10.30 UT in Fig. 5.
Parameter
a (km)
b (km)
c (km)
SEPβ (◦ )
PAnode (◦ )
ψ0 (UT)
EQJ2000 (α0 ,δ0 in ◦ )
σ radius (◦ )
ECJ2000 (λ0 ,β0 in ◦ )
Mean
379 ± 1
330 ± 1
225 ± 9
-41 ± 5
212 ± 1
9.74 ± 0.01
258;+11
1
256;+34
Table 3: Triaxial Ellipsoid Fit Parameters from 2007 observations. Uncertain-
ties reported here are formal error bars of the fit, see the tex t for a discussion on
the systematics.
10
11
12
13
14
15
12−Oct−2003
400
350
300
250
)
m
k
(
r
e
t
e
m
a
i
D
0.25
0.2
c
e
S
c
r
A
0.15
200
−90
−45
0
45
90
Rotational Phase
135
180
225
270
10
250
11
12
13
14
15
Conv
Deconv
Edges
s
i
x
A
g
n
o
L
f
o
)
°
(
A
P
220
190
160
−90
−45
0
45
90
Rotational Phase
135
180
225
270
Figure 8: Global fit and 2003 data.
3
14
17
10
11
15
16
12
13
6
7
8
9
20
4
5
1
2
19
18
the three parameters that differ due to the changing angles for
each date.
Statistical uncertainties for the dimensional parameters, as
well as those involving angles, such as pole position and lon-
gitude of the node, come from the non-linear least-squares fi t
for the 6 parameters that de fine the TE model, the 3 diameters
and 3 Euler angles. Systematic effects can arise in the process
of constructing a 3-D description of an asteroid from informa-
tion limited to a 2-D plane (images). Therefore, one needs to be
particularly vigilant regarding model assumptions, and their ap-
propriateness for a particular situation. While the uncertainties
derived for the parameters as fit by the model are straightfor -
ward, estimating the systematic effects that are present is not.
Deriving realistic (and therefore, directly applicable by other
workers) uncertainties for our results, including possible sys-
tematics, is the most challenging aspect of our work.
We have carefully calibrated some of these uncertainties by
making observations of external sources (e.g., the moons of
Saturn) of known size. One of the results of that work has
shown that our systematic uncertainties are larger for objects
of smaller angular diameter, until we reach a limit (at about
0.09′′ for a 10 m telescope) where we can no longer get reliable
sizes. Aspect ratios of projected shapes are still possible, but
absolute sizes break down. We have found that our systematics
from these tests span about 1 –4% per linear dimension. In ad-
dition, we have also imaged targets of spacecraft missions prior
to flyby (see KOALA section). In the case of (21) Lutetia, de-
spite an angular size of only 0.10′′ , our resulting models were
28
good to 2% in size and 2 km RMS in topography on a 100 km 29
object (see Carry et al. 2012).
30
We can also compare our TE results with those of KOALA
(see below), in cases where we have adequate observations. In
particular, we have such comparisons for four asteroids, includ-
ing (52) Europa. We can look for consistency, not only between
the two techniques, but in sub-sets of data to learn how far we
fall from the “correct” values. We can also compare the resul
ts
of data sets from different years. Our upcoming article, men-
tioned above (Drummond et al., in preparation) will be a stand-
alone treatment of the global fitting technique and calibrat ion
that will include much detail on uncertainties. For the present
results, we have determined that we should add quadratically
systematic uncertainties of 4.1%, 2.3%, and 3.8% to the TE-
derived fit errors (given in Table 4) for a, b, c, respectively. The
resulting total uncertainty estimates for the a, b, c diameters are
16 × 8 × 10 km, with a 7 degree systematic uncertainty for the
orientation of the spin axis. See Fig. 4 for a visualization of the
orientation of (52) Europa on the plane of the sky.
21
26
27
34
35
42
43
31
32
39
40
23
24
36
37
44
45
46
47
22
25
33
38
41
4. Comparison of (52) Europa to Lightcurves Inversion
Model
From optical lightcurves of (52) Europa, Micha łowski et al.
(2004) found a rotational pole at [λ0 , β0 ]=[252◦ , +38◦ ], with a
5◦ uncertainty in each Ecliptic coordinate. It is the pole closest
to ours in Fig. 6, about 6◦ away. They derived an a/b axial ra-
tio of 1.15, the same as our 1.15 ± 0.04, and a b/c ratio of 1.3,
compared with our 1.33 ± 0.05.
48
49
50
51
52
1
2
3
6
Diameter (km)
2003 Oct 12
2005 Jan 20
2007 May 28
a = 379 ± 2
+49 ± 1
+23 ± 1
-40 ± 1
b = 330 ± 2
204 ± 1
339 ± 1
213 ± 1
c = 249 ± 3
11.11 ± 0.02
10.31 ± 0.02
9.72 ± 0.02
Table 4: Results for the global fit. Uncertainties reported h ere are formal error bars of the fit.
Including systematic e ffects raises the total uncertainties to
16 × 8 × 10 km for the three ellipsoid diameters, and to 7◦ in the pole.
Pole
(α0 ,δ0 ) = 257◦;+12◦
σ radius = 1◦
(λ0 ,β0 ) = 255◦;+35◦
Param
SEPβ (◦ )
PAnode (◦ )
ψ0 (UT)
Figure 9: Comparison of our (2005) deconvolved K images from Fig. 3
(columns 1 and 3) with the web model of Michałowski et al. (2004), projected
forward from 1983 using their sidereal period of 0.2345816 days and an update
(although nearly identical) to thier pole from DAMIT (columns 2 and 4).
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Figure 9 is a side by side comparison of our decon-
volved Kp images, from 2005 January 20 (from Fig. 3) and
the Micha łowski et al. model, using the updated rotational
pole for the model at [λ0 , β0 ]=[251◦; +35◦ ] from the DAMIT
( Durech et al. 2010) web site1 . Figure 10 shows
compari-
son between our convolved and deconvolved images and the
lightcurves inversion model for 2003 and 2007.
The overall agreement between our AO deconvolved images
and the model predictions is excellent. A careful examina-
tion of Figs 9 and 10, however, will show edge features that
are seen in one but not the other, requiring the development
of an updated shape model, as discussed in following section.
Despite these features, (52) Europa is still well-modeled as a
smooth triaxial ellipsoid.
5. KOALA 3-D shape model
We construct a 3-D shape model of (52) Europa to give
a better rendering of the apparent shape visible in the im-
ages. For that, we use our KOALA algorithm (Carry et al.
2010a; Kaasalainen 2011) that makes combined use of optical
1 http://astro.troja.mff.cuni.cz/projects/asteroids3D/web.php
1
7
Figure 10: Same as Fig. 9, but for 2003 and 2007. In addition to the decon-
volved images in the middle row, we show the non-deconvoled, shifted, and
centered images in the top row for each epoch. In 2003, (52) Europa was 1.3
times closer than in 2007 resulting in different scales for the two years.
lightcurves, stellar occultations timings, and pro files fr om disk-
resolved images. The results of KOALA have been recently val-
idated at (21) Lutetia by the images taken by the ESA Rosetta
mission: The 3-D shape model and spin orientation determined
before the encounter by combining AO images and lightcurves
(Carry et al. 2010b) were in complete agreement with images
and results from the flyby ( Sierks et al. 2011; Carry et al. 2012).
Axial dimensions from KOALA were determined within 2% of
the the actual values and RMS differences in topography were
only 2 km.
We use here the 18 imaging epochs described in Sect. 2,
together with 49 lightcurves taken between 1979 and 2011
(we acquired 8 additional lightcurves within the CdR/CdL
collaboration with respect to the 41 lightcurves presented by
Micha łowski et al. 2004 ), and 4 stellar occultations (timings
taken from Dunham et al. 2011). A comparison of the KOALA
3-D shape model with the AO images from 2005 is presented
in Fig. 11. The agreement between the 3-D shape model and
the data is very good. The typical deviation with the 18 imag-
ing contours is of 0.2 pixel, corresponding to a few km. The
49 lightcurves are rendered at a level of 0.03 mag, i.e., close to
the intrinsic level of uncertainty of the data. Finally, the residu-
als between the occultation chords and the model are 13 km, on
22
average, mainly owing to the lower quality of 1983 occultation
timings (residuals of 19 km, compared to 11, 13, and 6 km for
the other epochs). Figure 12 shows these chords mapped onto
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
the projections of the 3-D KOALA model for the epochs of the
occultations.
The 3-D shape derived with KOALA is close to an ellipsoid,
validating (52) Europa as a Standard Triaxial Ellipsoid Asteroid
(STEA, Drummond et al. 2008). Fitting the KOALA model as
a triaxial ellipsoid yields diameters of 368 × 327 × 255 km, in
excellent agreement with the diameters and total uncertainties
in Table 4. The volume-equivalent spherical-diameter of the
KOALA 3-D shape model, derived by summing volume cells,
is 312 ± 6 km, in excellent agreement with the TE analysis pre-
sented above. The KOALA model yields a spin pole within
3◦of [λ0 , β0 ]=[254◦; +37◦ ] or [α0 , δ0 ]=[257◦; +15◦ ], also close
to the TE result. The shape model can be downloaded from
the DAMIT web page.
6. Occurrence of large facets on C-type asteroids
The 3-D shape model presents two broad shallow depres-
sions, probably best noted in the lower right of Fig. 12. They
can also be seen on the tops and bottoms of the asteroid im-
ages in column 1, panel 3, and column 3, panel 2. The depar-
tures from an ellipsoid, however, are not nearly as significa nt
as the apparent giant facets seen in our analysis of (511) Davida
(Conrad et al. 2007), nor as prominent, relative to body size, as
the giant craters seen on (253) Mathilde (Veverka et al. 1997).
We chose Mathilde as a prototypical object displaying giant
features seen in pro file (craters /facets), although Mathilde was
a much smaller asteroid than Davida. But it turns out that
(52) Europa is almost a twin of Davida in many respects: both
are C-type asteroids of very nearly the same size, they have
similar spin periods, and they have similar orbital properties,
so they have likely seen the same impact environment (al-
though Davida does have a bit larger eccenticity and inclina-
tion). In the Davida paper, we went to some length to demon-
strate that Davida could have encountered impacts of the size
necessary to produce the giant facets seen, without having bro-
ken up the body. So given the similarities between Davida
and (52) Europa, one might now wonder how likely it is that
(52) Europa would not show such facets (or at least not show
facets that are quite as prominent).
Returning to our analysis in the Davida paper, we estimated
that Davida should have had about 2.5 impacts large enough to
make such a giant crater during its lifetime. This led to the con-
clusion that if the facets seen were indeed craters, seen edge-on,
as on Mathilde, they would not be unexpected. The same statis-
tics should hold true for (52) Europa. But with an expected
total of only 2.5 impacts of this size during its lifetime, the
chances are also reasonable that it did not encounter any such
impacts. We therefore conclude that not seeing such prominent
features on a twin such as (52) Europa could also be expected.
Of course, the flux of smaller impactors would be higher, and
these would be responsible for the perhaps less prominent edge
features that we do see. Given that the viewing geometry has to
be just right to see these types of facets, it is possible that ob-
servational circumstances have conspired such that we missed
some giant feature, or such that those facets we do see are less
pronounced or are particularly hard to follow with rotation. We
Figure 11: Comparison of our 2005 deconvolved Kp images from Fig. 3
(columns 1 and 3) with the KOALA model described here
Figure 12: Comparison of the four stellar occultations with the KOALA shape
model. Solid and dashed grey lines represent positive (hits) and negative
(misses) chords, respectively. Black contours are the projection of the KOALA
3-D shape model on the plane of the sky at each occultation epoch.
have a fairly wide range of latitudes and longitudes in our data
set, however, so the chances of missing something as promi-
54
nent as a Davida-style facet are diminished, and we assert that
55
Europa appears qualitatively different than Davida.
56
57
58
59
1
8
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
7. Density of (52) Europa
There are 17 estimates of the mass of (52) Europa avail-
able in the literature, derived either from the analysis of the
orbit’s de flection during close approaches of minor planets
to
(52) Europa (e.g., Michalak 2001), or from a general adjust-
ment of the parameters used to generate the ephemeris of the
planets and asteroids in the Solar System (e.g., Fienga et al.
2009). We adopt here the weighted mean of these deter-
minations (following the selection discussed in Carry 2012):
(2.38 ± 0.58) × 1019 kg.
In general, the differences in volume between the triaxial and
the KOALA models are small. Here, that difference is less than
1%, which would lead to a volume difference of less than 3%.
When assigning uncertainties to our sizes (from either method),
we not only assess the derivable statistical uncertainties, but we
must also provide an estimate of systematic effects, of which
this difference is an example. The uncertainties used already
include potential differences between the models. Because of
the added topographic detail provided by the KOALA model,
we choose, in this case, to use the KOALA-derived volume of
(1.59 ± 0.10) ×107 km3 , giving a density of 1.5 ± 0.4 g cm−3 .
This bulk density falls within the observed range of densities
for C-type asteroids. Here, the uncertainty is mainly due to the
uncertainty on the mass determination (24%) rather than the
volume uncertainty of 6%. Thus, we are at the point in the study
of the density of asteroids where the uncertainty on the volume
is no longer the limiting factor (volume determination remains
generally the limiting factor when the mass is estimated from
a spacecraft encounter or a satellite, see the review by Carry
2012).
Dedicated observing programs and theoretical work are now
needed to derive more accurate masses of large main-belt aster-
oids. The advent of the Gaia mission (expected launch 2013)
should contribute a large number of new, improved masses (see
Mouret et al. 2007, for instance). With these more reliable vol-
umes and masses, we can derive improved densities and porosi-
ties, which in turn will allow us to better understand how den-
sity and porosity may be related to taxonomic class, absolute
diameters, or location (e.g., inner vs. outer main belt). And
this highlights the importance of continuing to push for more
AO observations of asteroids for size/shape determination, from
the best facilities, and the continued development of techniques,
such as KOALA, that combine multiple data types (hopefully,
eventually to include thermal radiometry and radar echoes).
8. Summary
At this point, (52) Europa can be considered for member-
ship as a Standard Triaxial Ellipsoid Asteroid (STEAs, see
Drummond et al. 2008) because it is so well modeled as an
ellipsoid (like asteroid (511) Davida, see Conrad et al. 2007).
The ellipses projected by these standard ellipsoids can be pre-
dicted well into the future or past, and therefore, can be used
as calibration objects for other techniques used in studying as-
teroids. Conrad et al. (2007) and Drummond et al. (in prepara-
tion) detail the equations necessary to predict the asterocentric
latitudes and longitudes, and Drummond (2000) show how to
derive the projected ellipse parameters from the asterocentric
latitudes and longitudes. For example, (52) Europa’s asterocen-
tric latitude can be predicted to within the error of its rotational
pole, 7◦ , and its asterocentric longitude to within 0.5◦ /yr since
the date of the most recent epoch reported here (2007 May 28).
The longitude uncertainty arises from the formal uncertainty in
the sidereal period, but in fact, judging by the good agreement
shown between the images and lightcurves inversion model pro-
jected forward from 1983, longitudes should be predictable to a
much higher accuracy than these values indicate. The projected
major or minor axis dimensions can be predicted to within ap-
proximately the uncertainty found here of 5 –10 km, and the or i-
entation of the apparent ellipse to within 2◦ .
We are fortunate to have both the triaxial ellipsoid (TE,
Drummond et al. 2009a) and the KOALA (Carry et al. 2010a)
techniques available for our analysis of AO images of aster-
oids. Each has its own strengths. TE requires relatively few im-
ages, can return shape/size/pole information amazingly quickly,
is generally insenstive to changes in the PSF, and is usually ade-
quate to get the basic asteroid parameters. For more detailed 3-
D shape information we can rely on KOALA. Unlike lightcurve
inversion alone, KOALA can obtain absolute sizes, and is sen-
sitive to concavities. The methods can be used to validate each
other, as we found exceedingly useful during our analysis of the
Lutetia data, prior to the Rosetta flyby ( Drummond et al. 2010;
Carry et al. 2010b). And while a detailed 3-D shape model
might be seen to supercede the triaxial assumption of TE, that
is not necessarily the case. As an example, our AO imaging of
the close flyby of Near-Earth Asteroid 2005 YU55 from Keck,
in November 2011, resulted in almost immediate size and shape
information from TE (Merline et al. 2011). In futher analysis,
we had hoped to use numerous lightcurves, taken near the time
of the flyby, to help re fine the size
/shape with KOALA. But de-
spite our efforts, the lightcurve information on 2005 YU55 so-
far is insufficient (mostly due to a very slow spin period) to al-
low KOALA to improve significantly on TE. This demonstrates
the importance of having both methods available for analysis of
our asteroid data.
New imaging, lightcurve, and occultation data will be added
to our overall analysis for (52) Europa as they become avail-
able. These may allow us to distinguish whether any of the
somewhat- flattened edges seen on (52) Europa in our existing
data sets are indeed facets or craters of the type seen on Davida
and Mathilde, and to better evaluate the extent and morphology
of any departure from a pure ellipsoid. The techniques we are
developing here (both observational and in data analysis) will
allow us to make immediate and substantial advances once data
from new, larger telescopes can be acquired.
Acknowledgments
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
This work presented here was supported by grants from 106
NASA’s Planetary Astronomy Program and the U.S. National
107
Science Foundation, Planetary Astronomy Program. We are
108
grateful for telescope time made available to us by the NASA 109
TAC, and also for the support of our collaborators on Team
1
9
Keck, the Keck science staff. The work of J. D was supported
by grant P209/10/0537 of the Czech Science Foundation and
by the Research Program MSM0021620860 of the Ministry of
Education.
This research has made use of NASA’s Astrophysics Data
System. The authors wish to recognize and acknowledge the
very significant cultural role and reverence that the summit of
Mauna Kea has always had within the indigenous Hawaiian
community. We are most fortunate to have the opportunity to
conduct observations from this mountain.
68
69
70
71
References
B. A. Archinal, M. F. A’Hearn, E. Bowell, A. Conrad, G. J. Consolmagno,
R. Courtin, T. Fukushima, D. Hestroffer, J. L. Hilton, G. A. Krasinsky,
G. Neumann, J. Oberst, P. K. Seidelmann, P. Stooke, D. J. Tholen, P. C.
Thomas, and I. P. Williams, 2011. Report of the IAU Working Group on Car-
tographic Coordinates and Rotational Elements: 2009. Celestial Mechanics
and Dynamical Astronomy , 109:101–135.
D. T. Britt, G. J. Consolmagno, and W. J. Merline. Small Body Density and
Porosity: New Data, New Insights. In S. Mackwell & E. Stansbery, editor,
37th Annual Lunar and Planetary Science Conference , volume 37 of Lunar
and Planetary Inst. Technical Report , page 2214, 2006.
D. T. Britt, D. K. Yeomans, K. R. Housen, and G. J. Consolmagno, 2002. As-
teroid Density, Porosity, and Structure. Asteroids III, pages 485–500.
B. Carry. Asteroids physical properties from high angular-resolution imaging .
PhD thesis, Observatoire de Paris, 2009.
B. Carry, 2012. Density of asteroids. Planetary and Space Science , 73:98–118.
B. Carry, C. Dumas, M. Fulchignoni, W. J. Merline, J. Berthier, D. Hestrof-
fer, T. Fusco, and P. Tamblyn, 2008. Near-Infrared Mapping and Physical
Properties of the Dwarf-Planet Ceres. Astronomy and Astrophysics , 478(4):
235–244.
B. Carry, C. Dumas, M. Kaasalainen, J. Berthier, W. J. Merline, S. Erard, A. R.
Conrad, J. D. Drummond, D. Hestroffer, M. Fulchignoni, and T. Fusco,
2010a. Physical properties of (2) Pallas. Icarus , 205:460–472.
B. Carry, M. Kaasalainen, C. Leyrat, W. J. Merline, J. D. Drummond, A. R.
Conrad, H. A. Weaver, P. M. Tamblyn, C. R. Chapman, C. Dumas, F. Co-
las, J. C. Christou, E. Dotto, D. Perna, S. Fornasier, L. Bernasconi,
R. Behrend, F. Vachier, A. Kryszczynska, M. Polinska, M. Fulchignoni,
R. Roy, R. Naves, R. Poncy, and P. Wiggins, 2010b. Physical properties
of the ESA Rosetta target asteroid (21) Lutetia. II. Shape and flyby geome-
try. Astronomy and Astrophysics , 523:A94.
B. Carry, M. Kaasalainen, W. J. Merline, T. G. M uller, L. Jorda, J. D. Drum-
Durech, M. K uppers, A. R. Conrad,
mond, J. Berthier, L. O’Rourke, J.
C. Dumas, H. Sierks, and the OSIRIS TeamPSS, 2012. KOALA shape mod-
eling technique validated at (21) Lutetia by ESA Rosetta mission. Planetary
and Space Science , 66:200–212.
A. R. Conrad, C. Dumas, W. J. Merline, J. D. Drummond, R. D. Campbell,
R. W. Goodrich, D. Le Mignant, F. H. Chaffee, T. Fusco, S. H. Kwok, and
R. I. Knight, 2007. Direct measurement of the size, shape, and pole of
511 Davida with Keck AO in a single night. Icarus , 191(2):616–627.
J. D. Drummond. Measuring Asteroids with Adaptive Optics. In N. Ageorges
and C. Dainty, editors, Laser Guide Star Adaptive Optics for Astronomy ,
pages 243–262, 2000.
J. D. Drummond, J. C. Christou, and J. Nelson, 2009a. Triaxial ellipsoid dimen-
sions and poles of asteroids from AO observations at the Keck-II telescope.
Icarus , 202:147–159.
J. D. Drummond, A. Conrad, W. Merline, and B. Carry, 2009b. The Di-
mensions and Pole of Asteroid (21) Lutetia from Adaptive Optics Images.
AAS/Division for Planetary Sciences Meeting Abstracts , 41:# 59.07.
J. D. Drummond, A. Conrad, W. J. Merline, B. Carry, C. R. Chapman, H. A.
Weaver, P. M. Tamblyn, J. C. Christou, and C. Dumas, 2010. Physical prop-
erties of the ESA Rosetta target asteroid (21) Lutetia. I. The triaxial ellipsoid
dimensions, rotational pole, and bulk density. Astronomy and Astrophysics ,
523:A93.
J. D. Drummond, R. Q. Fugate, J. C. Christou, and E. K. Hege, 1998. Full
Adaptive Optics Images of Asteroids Ceres and Vesta; Rotational Poles and
Triaxial Ellipsoid Dimensions. Icarus , 132:80–99.
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
J. D. Drummond, W. J. Merline, A. Conrad, C. Dumas, and B. Carry. Standard
Triaxial Ellipsoid Asteroids from AO Observations.
In AAS/Division for
Planetary Sciences Meeting Abstracts #40 , volume 40 of Bulletin of the
American Astronomical Society , page 427, 2008.
D. W. Dunham, D. Herald, E. Frappa, T. Hayamizu, J. Talbot, and B. Timerson.
Asteroid Occultations. NASA Planetary Data System, 2011. EAR-A-3-
RDR-OCCULTATIONS-V9.0.
J. Durech, V. Sidorin, and M. Kaasalainen, 2010. DAMIT: a database of aster-
oid models. Astronomy and Astrophysics , 513:A46.
A. Fienga, J. Laskar, T. Morley, H. Manche, P. Kuchynka, C. Le Poncin-Lafitte,
F. Budnik, M. Gastineau, and L. Somenzi, 2009. INPOP08, a 4-D planetary
ephemeris: from asteroid and time-scale computations to ESA Mars Express
and Venus Express contributions. Astronomy and Astrophysics , 507:1675–
1686.
M. Kaasalainen, 2011. Maximum compatibility estimates and shape recon-
struction with boundary curves and volumes of generalized projections.
Inverse Problems and Imaging , 5(1):37–57.
Q. M. Konopacky, A. M. Ghez, G. Duch ene, C. McCabe, and B. A. M acintosh,
2007. Measuring the Mass of a Pre-Main-Sequence Binary Star through the
Orbit of TWA 5A. Astronomical Journal, 133:2008–2014.
W. J. Merline, J. D. Drummond, P. M. Tamblyn, B. Carry, C. Neyman, A. R.
Conrad, C. R. Chapman, J. C. Christou, C. Dumas, and B. L. Enke, 2011.
2005 YU 55. IAU Circular , 9242:1.
W. J. Merline, J. D. Drummond, P. M. Tamblyn, B. Carry, C. Neyman, A. R.
Conrad, C. R. Chapman, J. C. Christou, C. Dumas, and B. L. Enke. Keck
Adaptive-Optics Imaging of Near-Earth Asteroid 2005 YU55 during its
In Asteroids, Comets, Meteors Meeting, Japan , num-
2011 Close Flyby.
ber 6372, 2012.
W. J. Merline, S. J. Weidenschilling, D. D. Durda, J.-L. Margot, P. Pravec, and
A. D. Storrs, 2002. Asteroids Do Have Satellites. Asteroids III, pages 289–
312.
G. Michalak, 2001. Determination of asteroid masses. II. (6) Hebe, (10) Hygiea,
(15) Eunomia, (52) Europa, (88) Thisbe, (444) Typtis, (511) Davida and
(704) Interamnia. Astronomy and Astrophysics , 374:703–711.
T. Michałowski, T. Kwiatkowski, M. Kaasalainen, W. Pych, A. Kryszczy ´nska,
P. A. Dybczy ´nski, F. P. Velichko, A. Erikson, P. Denchev, S. Fauvaud, and
G. M. Szab ´o, 2004. Photometry and models of selected main be lt asteroids I.
52 Europa, 115 Thyra, and 382 Dodona. Astronomy and Astrophysics , 416:
353–366.
S. Mouret, D. Hestroffer, and F. Mignard, 2007. Asteroid masses and improve-
ment with GAIA. Astronomy and Astrophysics , 472:1017–1027.
L. M. Mugnier, T. Fusco, and J.-M. Conan, 2004. MISTRAL: a Myopic Edge-
Preserving Image Restoration Method, with Application to Astronomical
Adaptive-Optics-Corrected Long-Exposure Images. Journal of the Optical
Society of America A, 21(10):1841–1854.
D. Nesvorn ´y, W. F. Bottke, Jr., L. Dones, and H. F. Levison, 2 002. The recent
breakup of an asteroid in the main-belt region. Nature, 417:720–771.
H. Sierks, P. Lamy, C. Barbieri, D. Koschny, H. Rickman, R. Rodrigo, M. F.
A’Hearn, F. Angrilli, A. Barucci, J.-L. Bertaux, I. Bertini, S. Besse, B. Carry,
G. Cremonese, V. Da Deppo, B. Davidsson, S. Debei, M. De Cecco,
J. De Leon, F. Ferri, S. Fornasier, M. Fulle, S. F. Hviid, G. W. Gaskell,
O. Groussin, P. J. Gutierrez, L. Jorda, M. Kaasalainen, H. U. Keller, J. Knol-
lenberg, J. R. Kramm, E. K uhrt, M. K uppers, L. M. Lara, M. Lazzarin,
C. Leyrat, J. L. Lopez Moreno, S. Magrin, S. Marchi, F. Marzari, M. Mas-
sironi, H. Michalik, R. Moissl, G. Naletto, F. Preusker, L. Sabau, W. Sabolo,
F. Scholten, C. Snodgrass, N. Thomas, C. Tubiana, P. Vernazza, J.-B. Vin-
cent, K.-P. Wenzel, T. Andert, M. P atzold, and B. P. Weiss, 2011. Images of
asteroid (21) Lutetia: A remnant planetesimal from the early Solar System.
Science , 334:487–490.
P. C. Thomas, R. P. Binzel, M. J. Gaffey, A. D. Storrs, E. N. Wells, and B. H.
Zellner, 1997. Impact excavation on asteroid 4 Vesta: Hubble Space Tele-
scope results. Science , 277:1492–1495.
M. A. van Dam, E. M. Johansson, P. J. Stomski Jr., R. Sumner, J. C. Y. Chin,
and P. L. Wizinowich. Performance of the Keck II AO system, 2007.
M. A. van Dam, D. Le Mignant, and B. Macintosh, 2004. Performance of the
Keck Observatory adaptive-optics system. Applied Optics , 43(23):5458–
5467.
J. Veverka, P. C. Thomas, A. Harch, B. E. Clark, J. F. Bell, B. Carcich, J. Joseph,
C. R. Chapman, W. J. Merline, M. S. Robinson, M. Malin, L. A. McFadden,
S. L. Murchie, S. E. Hawkins, R. Faquhar, N. Izenberg, and A. F. Cheng,
1997. NEARs Flyby of 253 Mathilde: Images of a C Asteroid. Science ,
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
603
604
605
606
607
608
609
610
611
612
613
614
615
10
278:2109–2114.
J. Veverka, P. C. Thomas, A. Harch, B. E. Clark, B. Carcich, J. Joseph, S. L.
Murchie, N. Izenberg, C. R. Chapman, W. J. Merline, M. Malin, L. A.
McFadden, and M. S. Robinson, 1999. NEAR Encounter with Asteroid
253 Mathilde: Overview. Icarus , 140:3–16.
616
617
618
619
620
11
|
1910.12899 | 1 | 1910 | 2019-10-28T18:28:22 | K2-19b and c are in a 3:2 Commensurability but out of Resonance: A Challenge to Planet Assembly by Convergent Migration | [
"astro-ph.EP"
] | K2-19b and c were among the first planets discovered by NASA's K2 mission and together stand in stark contrast with the physical and orbital properties of the solar system planets. The planets are between the size of Uranus and Saturn at 7.0$\pm$0.2 R_E and 4.1$\pm$0.2 R_E, respectively, and reside a mere 0.1% outside the nominal 3:2 mean-motion resonance. They represent a different outcome of the planet formation process than the solar system, as well as the vast majority of known exoplanets. We measured the physical and orbital properties of these planets using photometry from K2, Spitzer, and ground-based telescopes, along with radial velocities from Keck/HIRES. Through a joint photodynamical model, we found that the planets have moderate eccentricities of $e \approx0.20$ and well-aligned apsides $\Delta \varpi \approx 0$ deg. The planets occupy a strictly non-resonant configuration: the resonant angles circulate rather than librate. This defies the predictions of standard formation pathways that invoke convergent or divergent migration, both of which predict $\Delta \varpi \approx 180$ deg and eccentricities of a few percent or less. We measured masses of $M_{p,b}$ = 32.4$\pm$1.7 M_E and $M_{p,c}$ = 10.8$\pm$0.6 M_E. Our measurements, with 5% fractional uncertainties, are among the most precise of any sub-Jovian exoplanet. Mass and size reflect a planet's core/envelope structure. Despite having a relatively massive core of $M_{core} \approx15$ $M_E$, K2-19b is envelope-rich, with an envelope mass fraction of roughly 50%. This planet poses a challenge to standard models core-nucleated accretion, which predict that cores $\gtrsim 10$ $M_E$ will quickly accrete gas and trigger runaway accretion when the envelope mass exceeds that of the core. | astro-ph.EP | astro-ph | Draft version October 30, 2019
Typeset using LATEX twocolumn style in AASTeX63
K2-19b and c are in a 3:2 Commensurability but out of Resonance:
A Challenge to Planet Assembly by Convergent Migration
Erik A. Petigura,1 John Livingston,2, ∗ Konstantin Batygin,3 Sean M. Mills,4 Michael Werner,5
Howard Isaacson,6, 7 Benjamin J. Fulton,8 Andrew W. Howard,4 Lauren M. Weiss,9, † Néstor Espinoza,10, ‡
Daniel Jontof-Hutter,11 Avi Shporer,12 Daniel Bayliss,13, § and S. C. C. Barros13
9
1
0
2
t
c
O
8
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
9
8
2
1
.
0
1
9
1
:
v
i
X
r
a
1Department of Physics & Astronomy, University of California Los Angeles, Los Angeles, CA 90095, USA
2Department of Astronomy, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan
3Division of Geological and Planetary Sciences, California Institute of Technology, Pasadena CA, 91125, USA
4Cahill Center for Astrophysics, California Institute of Technology, Pasadena CA, 91125, USA
5Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA
6Department of Astronomy, University of California Berkeley, Berkeley CA 94720
7University of Southern Queensland, Toowoomba, QLD 4350, Australia
8IPAC-NASA Exoplanet Science Institute Pasadena, CA 91125, USA
9Institute for Astronomy, University of Hawaii at Manoa, Honolulu, HI 96822, USA
10Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidelberg, Germany
11Department of Physics, University of the Pacific, Stockton, CA 95211, USA
12Department of Physics and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, Cambridge, MA
02139, USA
13Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, PT4150-762 Porto, Portugal
ABSTRACT
K2-19b and c were among the first planets discovered by NASA's K2 mission and together stand in
stark contrast with the physical and orbital properties of the solar system planets. The planets are
between the size of Uranus and Saturn at 7.0± 0.2 R⊕ and 4.1± 0.2 R⊕, respectively, and reside a mere
0.1% outside the nominal 3:2 mean-motion resonance. They represent a different outcome of the planet
formation process than the solar system, as well as the vast majority of known exoplanets. We measured
the physical and orbital properties of these planets using photometry from K2 , Spitzer, and ground-
based telescopes, along with radial velocities from Keck/HIRES. Through a joint photodynamical
model, we found that the planets have moderate eccentricities of e ≈ 0.20 and well-aligned apsides
∆ ≈ 0 deg. The planets occupy a strictly non-resonant configuration: the resonant angles circulate
rather than librate. This defies the predictions of standard formation pathways that invoke convergent
or divergent migration, both of which predict ∆ ≈ 180 deg and eccentricities of a few percent or less.
We measured masses of Mp,b = 32.4 ± 1.7 M⊕ and Mp,c = 10.8 ± 0.6 M⊕. Our measurements, with
5% fractional uncertainties, are among the most precise of any sub-Jovian exoplanet. Mass and size
reflect a planet's core/envelope structure. Despite having a relatively massive core of Mcore ≈ 15 M⊕,
K2-19b is envelope-rich, with an envelope mass fraction of roughly 50%. This planet poses a challenge
to standard models core-nucleated accretion, which predict that cores (cid:38) 10 M⊕ will quickly accrete
gas and trigger runaway accretion when the envelope mass exceeds that of the core.
Keywords: planets and satellites: individual (K2-19b,K2-19c) -- planets and satellites: dynamical evo-
formation -- techniques: radial velocities --
lution and stability -- planets and satellites:
techniques: photometric
Corresponding author: Erik A. Petigura
[email protected]
∗ JSPS Fellow
† Parrent Postdoctoral Fellow
‡ Bernoulli Fellow
IAU-Gruber Fellow
2
1. INTRODUCTION
While a perennial quest in exoplanet astronomy is the
discovery and characterization of ever more "Earth-like"
worlds, our understanding of planet formation is best
informed by the full diversity of planets around other
stars. Thanks to the rapidly growing census of extraso-
lar planets, we may now study the diverse outcomes of
planet formation processes beyond those that occurred
in the solar system. The K2-19 system is one such out-
come.
The system hosts three known planets. Armstrong
et al. (2015) initially reported K2-19b and c based on
photometry collected by the Kepler Space Telescope op-
erating in its K2 mode (Howell et al. 2014). K2-19b has
an orbital period of 7.9 days and has a radius of 7.0 R⊕,
between the size of Uranus and Saturn. K2-19c has an
orbital period of 11.9 days and a radius of 4.1 R⊕. While
K2-19c is similar in size to the solar system ice giants,
aspects of its bulk composition, such as ice fraction, may
be quite different due its close-in orbit. As techniques
to correct for K2 systematics improved, Sinukoff et al.
(2016) detected a third planet, K2-19d, a 1.2 R⊕ planet
on 2.5 day orbit.
In this paper, we focus on K2-19b and c, which reside
just outside the nominal 3:2 mean-motion resonance.
While Armstrong et al. (2015) detected transit-timing
variations (TTVs) within the K2 dataset, the relatively
short 80 day baseline resulted in significant uncertainties
in the TTV model. Several groups have subsequently
observed transits of K2-19b from the ground in order to
better constrain the TTV model (Armstrong et al. 2015;
Narita et al. 2015; Barros et al. 2015). However, to date,
there have been no successful recoveries of the K2-19c
transit, which has contributed to lingering uncertainty
in the TTV solution.
In parallel, several groups have obtained radial ve-
locity (RV) measurements of K2-19 in order to directly
constrain the planet masses through stellar reflex motion
(Dai et al. 2016; Nespral et al. 2017). A key challenge to
these efforts is that at V = 13.0 mag, K2-19 is near the
faint limit of most current RV facilities. In addition, the
star exhibits significant RV variability due to spot mod-
ulation, which must be disentangled from the planetary
signals.
In this work, we present the results of a coordinated
observational campaign to characterize K2-19b and c,
using both TTVs and RVs. We describe our photometry
in Section 2 and our RVs in Section 3. Our photomet-
ric dataset includes two Spitzer observations for each
planet. Our Spitzer observations of the K2-19c transits
are significant in that they are the first K2 and help to
reduce uncertainties in the TTV solution. We perform
a photodynamical analysis in Section 4, which yields
the most precise constraints on the masses and orbits
of these two planets to date. In Sections 5 -- 7 we assess
the bulk composition of these planets, their dynamical
evolution, and possible formation pathways.
2. PHOTOMETRIC OBSERVATIONS
2.1. K2 Photometry
The Kepler Space Telescope observed K2-19 from
2014-05-30 to 2014-8-21 during campaign 1 of its K2
mission. The photometry contain large systematics due
to pointing drifts of ∼1 pixel that occur on ∼6 hr
timescales. We used the EVEREST2.0 package to cor-
rect for these systematics (Luger et al. 2017), and the
corrected light curve is shown in Figure 1.
There is clear periodic variability with P ≈ 20 days
with a peak-to-trough amplitude of 1% due to rotation-
induced spot modulation. Figure 2 is a zoomed in view
of individual transits, some of which are overlapping.
2.2. Spitzer Photometry
The K2 data alone samples only a small fraction of
the multi-year TTV signal. We used the Spitzer Space
Telescope to observe two additional transits of K2-19b
and K2-19c to better sample this signal. Planet b obser-
vations were conducted on 2017-04-23 and 2017-09-05;
planet c observations were conducted on 2016-10-04 and
2017-04-08.1
To plan the first set of Spitzer observations, we con-
sulted the transit times predicted by Barros et al. (2015)
(S. Barros, private communication). Because our first
Spitzer observation of K2-19b was two years after the
last transit used in the Barros et al. (2015) model, there
was considerable timing uncertainty. We observed for 12
hours to reliably catch the 3.5 hour transit. There was
even more timing uncertainty for K2-19c, which had not
been observed since 2014, and we scheduled a 27 hour
observing sequence.
When planning our second set of Spitzer observations,
we constructed a preliminary TTV model with plausi-
ble values for the planet masses and eccentricity. Having
incorporated the first set of observations, there was less
uncertainty in the transit times of K2-19b and c, requir-
ing only 7 and 9 hour observing sequences, respectively.
§ Observatoire Astronomique de l'Université de Genève, 51 ch.
des Maillettes, 1290 Versoix, Switzerland
1 All observations were carried out under GO program 13052 (PI:
M. Werner).
3
Figure 1. Photometry from K2 after removing instrument systematics showing ≈ 1% periodic variability with P ≈ 20 days
(see Section 2.1). The red line is our Gaussian Process fit to the photometry, which informs the adopted noise model in our RV
analysis (see Section 4.1).
Figure 2. The black circles show the detrended K2 photometry around the transits. Several overlapping transits are observed.
The maximum a posteriori model is shown as the orange line and the residuals to this model are shown below. Increased scatter
during transit due to spot crossing events are observed during some transits (see Section 4.2).
198019902000201020202030204020502060BJD - 24548330.9900.9951.0001.005Normalized FluxphotometryGP0.9951.000Flux−0.0010.0000.001Resid0.9951.000Flux−0.0010.0000.001Resid0.9951.000Flux−0.20.00.2Time(days)−0.0010.0000.001Resid−0.20.00.2Time(days)−0.20.00.2Time(days)−0.20.00.2Time(days)−0.20.00.2Time(days)4
We used IRAC channel 2 (4.5 µm) because the instru-
mental systematics due to intra-pixel sensitivity vari-
ations are smaller than in channel 1 (3.6 µm; Ingalls
et al. 2012). We used 2 second exposures to optimize
the integration efficiency while remaining in the linear
regime of the detector. We extracted photometry from
the Spitzer data using circular apertures. As described
in Livingston et al. (2019), we selected the aperture size
(r = 2.2 pixels) that minimized the combined uncorre-
lated (white) and correlated (red) noise, as measured by
the standard deviation and β factor (Pont et al. 2006;
Winn et al. 2008). We resampled the light curve into
60 second integrations which yields improved system-
atic modeling without significantly altering the transit
profile (Benneke et al. 2017).
Following standard practice, we modeled the Spitzer
systematics and transit profile simultaneously. Using
the pixel-level decorrelation (PLD) method of Deming
et al. (2015), we constructed our systematic model from
a linear combination of the nine pixel-level lightcurves
from a 3 × 3 pixel grid centered on the star. For K2-
19b and c, we modeled each set of two transits simul-
taneously and shared all transit parameters except for
the transit mid-times Tc,i. We used a quadratic limb-
darkening parameterization and physically motivated
priors (Claret et al. 2012; Kipping 2013). In summary,
our modeling of each planet involved 28 free parame-
ters: nine PLD coefficients for each dataset, a white
noise term for each dataset, two transit mid-points Tc,i,
the orbital period P , the planet-star radius ratio Rp/R(cid:63),
the scaled semi-major axis a/R(cid:63), the impact parameter
b, the limb-darkening parameters q1 and q2.
We explored the range of coefficients allowed by our
data using the affine-invariant Markov Chain Monte
Carlo (MCMC) sampler of Goodman & Weare (2010).
We initialized 100 walkers and allowed them to evolve for
Nsteps = 10000 steps. We visually inspected the trace
plots and discarded the first 5000 steps of burn-in. We
assessed the convergence by computing the autocorrela-
tion length τ for each chain. We computed the mean
value of τ for all 100 chains for each parameter, and
found that Nsteps/τ ≥ 38 for all chains. Our corrected
light curves had an RMS scatter of ∼400 ppm on 40
minute timescales. The Spitzer photometry and best-fit
transit models are shown Figure 3. The derived transit
times are listed in Table 1.
2.3. Ground-based Photometry
We also included several transit times of K2-19b mea-
sured using ground-based facilities. Three were drawn
from Narita et al. (2015). We also observed a transit on
2017-06-05 with the 1m telescope of the Las Cumbres
Table 1. Transit Times
Planet Transit
Instrument
Tc
days
σ(Tc) Notes
30
34
41
133
87
141
150
102
FLWO
2218.0041
TRAPPIST 2249.6955
2305.1505
MuSCAT
3033.8604
Spitzer
3019.4774
Spitzer
LCO
3097.2502
3168.5368
Spitzer
Spitzer
3197.8645
K2-19b
K2-19b
K2-19b
K2-19b
K2-19c
K2-19b
K2-19b
K2-19c
Note -- Following a convention from the Kepler mission, times are
given in BJDTBD − 2454833. Notes -- A: This work; B: Narita
et al. (2015)
B
B
B
A
A
A
A
A
days
0.0022
0.0014
0.0014
0.0009
0.0074
0.0024
0.0014
0.0059
Observatory network (LCO; Brown et al. 2013), located
at the South African Astronomical Observatory. We
performed bias, dark, and flat-field corrections using the
standard LCOGT pipeline (McCully et al. 2018). We
then performed aperture photometry on K2-19 and 10
comparison stars having similar 2MASS colors and per-
formed differential photometry to remove instrumental
and atmospheric effects. We modeled the transit using
both white and correlated noise models and found that
the white-noise mode was preferred. The light curve and
transit fit are shown in Figure 4. The RMS scatter in
the residuals is ≈ 300 ppm per 40 min interval.
3. RADIAL VELOCITY OBSERVATIONS
We obtained 51 spectra of K2-19 using the High Res-
olution Echelle Spectrometer (HIRES; Vogt et al. 1994)
on the 10m Keck-I telescope between 2015-02-05 and
2017-12-26. We collected spectra through an iodine cell
mounted directly in front of the spectrometer slit. The
iodine cell imprints a dense forest of absorption lines
which serve as a wavelength reference. We also obtained
a "template" spectrum without iodine.
At V = 13.0 mag K2-19 is a challenging RV target for
Keck/HIRES. We aimed to achieve a consistent signal-
to-noise ratio (SNR) of 100 per reduced pixel at 5500 Å
using an exposure meter. However, various through-
put losses due to poor/variable seeing and cirrus clouds
sometimes resulted in lower than desired SNR. Our spec-
tra have per pixel SNR ranging from 53 to 108.
RVs were determined using standard procedures of the
California Planet Search (Howard et al. 2010) including
forward modeling of the stellar and iodine spectra con-
volved with the instrumental response (Marcy & Butler
1992; Valenti et al. 1995). The measurement uncertainty
of each RV point is derived from the uncertainty on the
5
Figure 3. Transits of K2-19b and c observed by Spitzer in the 4.5 µm IRAC channel along with our simultaneous modeling of
the instrumental systematics and transit profiles (see Section 2.2). Panel (a) shows the transit of K2-19b with a transit number
i = 133, where i = 0 corresponds to the first K2 transit. In the top sub-panel, we show the raw light curve (gray), the maximum
a posteriori (MAP) transit/systematic model (red), and the 95% credible models (light red band). In the bottom sub-panel, we
show the MAP corrected photometry (gray) and transit model (purple). The 95% credible models are shown with the purple
band. Panel (b): same as (a) but for the second Spitzer observation of K2-19b (i = 150). Panel (c): same as (a) but for the
first Spitzer observation of K2-19c (i = 87). Panel (d): same as (a) but for the second Spitzer observation of K2-19c (i = 102).
mean RV of the ∼700 spectral chunks used in the RV
pipeline and ranges from 1.9 to 3.8 m s−1. Table 2 lists
the RVs and uncertainties.
4. TTV AND RV MODELING
Here, we describe our modeling of both the photomet-
ric and RV datasets. In Section 4.1, we perform a Keple-
rian analysis of the RVs only. We observe quasiperiodic
RV variability due to rotating starspots, which we model
with a Gaussian process. Section 4.2 describes our pho-
todynamical analysis that incorporates constraints from
both photometry and RVs. This analysis yields tighter
constraints on the properties of K2-19b and c and the
parameters listed in Table 4.2 constitute our adopted
system parameters.
While the photodynamical analysis yields smaller un-
certainties, we present the RV-only analysis for the fol-
0.9900.9951.0001.0051.0101.015Normalized flux(a)transit+systematicsNormalized flux(b)transit+systematics3033.63033.73033.83033.93034.03034.1Time [BJD-2454833]0.9900.9951.0001.0051.0101.015Normalized fluxtransit3168.403168.453168.503168.553168.603168.653168.70Time [BJD-2454833]Normalized fluxtransit0.9900.9951.0001.0051.010Normalized flux(c)transit+systematicsNormalized flux(d)transit+systematics3019.23019.33019.43019.53019.63019.7Time [BJD-2454833]0.9900.9951.0001.0051.010Normalized fluxtransit3197.603197.653197.703197.753197.803197.853197.903197.95Time [BJD-2454833]Normalized fluxtransit6
Figure 4. Top panel: black points show the relative pho-
tometry of K2-19 observed by LCO on 2017-06-05 during the
transit of K2-19b (see Section 2.3). The red line is the best
fit transit model and the bottom panel shows the residuals
to the fit.
Table 2. Radial Velocities
Time
days
σ(RV)
RV
m s−1 m s−1
SHK
2.69
2.84
1.98
2.04
3.37
2.20
2.13
3.15
2.18
2.12
0.358
0.328
0.181
0.247
0.221
0.182
0.221
0.195
0.249
0.269
−8.92
2225.996346
2229.058283 −14.53
2346.849965 −11.35
−0.21
2366.792920
−9.39
2367.829151
2368.814357 −14.07
2370.809676 −18.62
6.22
2374.805352
2375.803685
2.94
2376.797458 −14.34
Note -- Radial velocities and uncertainties
for K2-19 (see Section 3). Times are
given in BJDTBD − 2454833. We also
provide the Mount Wilson SHK activ-
ity index (Vaughan et al. 1978), which
is measured to 1% precision. Table 2
is published in its entirety in machine-
readable format. A portion is shown here
for guidance regarding its form and con-
tent.
4.1. Keplerian RV modeling
We analyzed the RV timeseries using the open source
package RadVel (Fulton et al. 2018). RadVel facilitates
maximum a posteriori (MAP) model fitting and param-
eter estimation via MCMC. In general, a Keplerian RV
signal may be described by the orbital period P , time of
inferior conjunction Tc, eccentricity e, argument of peri-
astron ω, and Doppler semi-amplitude K. We included
K2-19b, c, and d in our model with zero eccentricity.
For planets b and c we fixed P and Tc to the mean
value as determined by the K2 and Spitzer photome-
try. For planet d, we fixed P and Tc to the Sinukoff
et al. (2016) ephemeris. While the planets do not have
strictly linear ephemerides, we confirmed that the errors
introduced by this simplification are negligible after per-
forming the photodynamical analysis described in Sec-
tion 4.2. In our preliminary fitting, we found that mod-
els with a linear acceleration term dv/dt were favored
by the Bayesian Information Criterion (BIC; Schwarz
1978) with ∆BIC = −15. In our subsequent modeling,
described below, we found dv/dt = 5.9±2.4 m s−1 yr−1.
The K2-19 photometry shows clear spot modulation
(see Figure 1), which can introduce correlated noise into
the RV timeseries. We estimated the amplitude of this
noise using the F F (cid:48) method of Aigrain et al. (2012):
∆RV∼ F F
(cid:48)
R(cid:63)/f
∼ (0.5%)(1%/5 d)(0.82 R(cid:12))/(1%)
∼ 7 ms
−1.
Here, F is the fractional flux variation, F (cid:48) is its time
derivative, and f is the maximum flux decrement due to
spots. This noise source is quasiperiodic as spots rotate
with the stellar photosphere and also evolve with time.
Numerous prior studies have modeled spot noise with
quasiperiodic Gaussian Processes (GPs) including Hay-
wood et al. (2014), Grunblatt et al. (2015), and others.
We used the following quasiperiodic kernel that specifies
the covariance between the i and j measurements:
(cid:20)
Ci,j = η2
+(cid:2)σ2
1exp
−
i + σ2
jit
(ti − tj)2
(cid:3) δi,j.
η2
2
1
2η2
4
−
sin2 π(ti − tj)
η2
3
(cid:21)
lowing reasons: (1) The RVs provide sensitivity to non-
transiting planets that could compromise the accuracy
of the photodynamical model. (2) RV variability from
rotating starspots is comparable in amplitude to that
due to K2-19b and may account for discrepancies be-
tween previously published mass measurements.
(3)
The two analyses demonstrate the relative strengths and
weaknesses of the TTV and RV techniques as probes of
the properties of the K2-19 system.
Here, η1 is the covariance amplitude, η2 is the expo-
nential decay length, η3 sets the period, η4 sets the
relative importance of the exponential decay part of
the kernel, and δi,j is the Kronecker delta function.
We trained the GP on the K2 photometry and found
η1 = 0.4 ± 0.1%, η2 = 34 ± 5 days, η3 = 20.4 ± 0.3 days,
and η4 = 0.49± 0.06. Our value for η3 is consistent with
our visual assessment of the stellar rotation period of
P ≈ 20 days.
0.9900.9951.0001.005Relative flux3097.203097.223097.243097.263097.283097.303097.323097.34Time (BJD - 2454833)500005000Residuals (ppm)We then modeled the RVs using the GP-based like-
lihood (see RadVel documentation for details). We
imposed Gaussian priors on η2, η3, and η4 based on
our photometric modeling described above.
In sum-
mary, our RV model had the following free parameters:
{K1, K2, K3, dv/dt, η1, η2, η3, η4, γ, σjit}.
Figure 5 shows the MAP model. We derived uncer-
tainties using MCMC, terminating the chains when the
inter-ensemble GR statistic was less than 1.003. For
K2-19b, we measured a mass of 33 ± 5 M⊕. The RVs
were insufficient to detect planaets c or d, but we placed
upper limits on their masses of Mp,c < 10.2 M⊕ and
Mp,d < 3.5 M⊕ at 95% confidence.
We found that η1, the amplitude of the quasiperi-
odic RV variability included in our GP noise model was
7.4 ± 2.2 m s−1, in agreement with our previous esti-
mate. This value is comparable to reflex velocity of
planet b, and it underscores the importance of treat-
ing spot-induced RV-variability in the RV analysis. We
recommend that future RV campaigns targeting K2-19
(or similar stars) observe at high cadence to better trace
this quasiperiodic noise source.
We explored fits where eb and ωb were allowed to vary.
However, this additional model complexity was disfa-
vored by the BIC, with ∆BIC = −5. Therefore, the RVs
alone are insufficient to detect eccentricity for K2-19b.
We characterized the values of eb excluded solely by the
RVs by running a second MCMC where √eb cos ωb and
√eb sin ωb were allowed to vary. We found that eb < 0.27
at 95% confidence, which is consistent with our photo-
dynamical analysis presented in Section 4.2.
We note that our RV-only mass measurement of planet
b is inconsistent at the ∼3σ level with that of Nespral
et al. (2017), who reported Mp,b = 54.8 ± 7.5 M⊕. The
Nespral et al. (2017) analysis used 22 RVs from three
different instruments: FIES, HARPS-N, and HARPS.
We hypothesize that, in the Nespral et al. (2017) analy-
sis, biases due to stellar activity were amplified given the
sparse sampling of the RV timeseries and offsets between
the RV datasets.
As we show in Section 4.2, the constraints from TTVs
on the masses and eccentricities of K2-19b and c are
more precise than those from the RVs. However, the
RVs provide sensitivity to non-transiting planets that
could compromise the accuracy of the TTV model. Non-
transiting planets near first order MMR are the most
concerning, as they would produce the largest TTVs.
To search for such planets, we computed the Lomb-
Scargle periodogram (Lomb 1976; Scargle 1982) of the
residuals to the most probable Keplerian model (see Fig-
ure 5). We found no additional signals with a boot-
strap false alarm probability of < 10% (VanderPlas
7
2018). Detection of an exoplanet from RVs alone with
P less than the observing baseline generally requires
K (cid:38) ασRV/√Nobs, where σRV is the individual RV mea-
surement uncertainty and α is a numerical prefactor of
≈10 (Howard & Fulton 2016). Adopting σRV = 9 m s−1,
the quadrature sum of the two dominant noise terms,
σjit and η1, we found that a planet with K (cid:38) 13 m s−1
would have been detectable. Therefore, at orbital peri-
ods comparable to those of K2-19b and c, the RVs rule
out planets with masses comparable to K2-19b. This
supports the assumption in our photodynamical model
that the TTV signal is dominated by interactions be-
tween K2-19b and K2-19c.
4.2. Photo-dynamical analysis
To extract the information contained in both the RV
and photometric datasets, we performed a photodynam-
ical analysis. We used the Phodymm code, which is
described in Mills et al. (2016). Given an initial config-
uration, Phodymm performs an N-body integration and
forward models the light curve. The forward modeling
approach has the advantage that it naturally handles si-
multaneous transits (Pál 2008) and simultaneously mod-
els all transit characteristics such as duration and depth
variations, compared to other techniques that model
derived transit times (see, e.g., TTVFast; Deck et al.
2014).
For each planet, we specified an initial set of osculat-
ing elements: P , Tc, e, ω, i, Ω. Here, i is the incli-
nation and Ω is the longitude of ascending node. The
model also requires Mp and Rp/R(cid:63) for each planet, and
the following stellar parameters: M(cid:63), R(cid:63), and quadratic
limb-darkening parameters, q1 and q2.
Because Ω is defined with respect to an arbitrary ref-
erence direction, we may fix Ωb to 0 deg without loss of
generality. K2-19d is dynamically decoupled from K2-
19b and c and does not significantly affect the transits
of the other planets gravitationally. However, K2-19d
sometimes transits at the same time as K2-19b or c
and therefore must be modeled out. We fixed ed = 0,
ωd = 0 deg, and Ωd = 0 deg. Following the recommen-
dations of Eastman et al. (2013), for planets b and c,
we parameterized {e, ω} as {√e cos ω,√e sin ω}, which
enforces a uniform prior on e. In total, our model had
24 free parameters.
To assess the degree to which our model fits the K2
photometry, we defined
(cid:88)
i
χ2
phot =
(cid:18) fmod,i − fi
(cid:19)2
,
σi
where fmod,i, fi, σi is the modeled flux, observed flux,
and flux uncertainty of the ith K2 observation.
8
Figure 5. The three-Keplerian fit to the K2-19 radial velocities (RVs), assuming circular orbits (see Section 4.1). Panel (a):
Points show RVs from HIRES and the line shows the most probable Keplerian model. The gray band shows Gaussian process
model that accounts for quasiperiodic correlated noise due to star spots. Panel (b) shows the phase-folded RVs and the most
probable Keplerian model for K2-19b with contributions from the GP noise model, dv/dt term, and other Keplerians removed.
Panel (c), same as (b), but for K2-19c. Panel (d), same as (b) but for K2-19d.
For the Spitzer and ground based transits, we modeled
the derived transit times (Table 1) rather than the pho-
tometry directly because it is impractical to marginalize
over the various systematic noise models that were used
to derive the transit times. We defined the following
goodness-of-fit statistic:
(cid:18) Tc,mod,j − Tc,j
(cid:19)2
(cid:88)
j
χ2
times =
σj
where Tc,mod,j, Tc,j, and σj are the modeled midpoint,
observed midpoint, and timing uncertainty of the jth
transit. Our final adopted log-likelihood is
log L = −
1
2
χ2
phot −
1
2
χ2
times.
Following Petigura et al. (2018a), we incorporated the
RV mass constraints as Gaussian priors on the planet
masses. We checked that this treatment is justified by
verifying that the posteriors on K1, K2, and K3 (Sec-
tion 4.1) are Gaussian and uncorrelated. Finally, we
applied Gaussian priors on M(cid:63) and R(cid:63) based on our
stellar characterization (see Table 4.2).
We explored the range of plausible models using Dif-
ferential Evolution Markov Chain Monte Carlo (DEM-
CMC). We ran 40 walkers and checked for convergence
by periodically computing the Gelman-Rubin (GR)
statistic (Gelman & Rubin 1992). We terminated our
runs after 80,000 steps, when the GR statistic was less
than 1.05 for all parameters. After inspecting the chains,
we discarded the first 10,000 steps as burn-in.
Figure 2 shows the MAP photodynamical fit to the
K2 dataset. We note that there is increased scatter
in the residuals during transits due to spot crossings
events. These spot crossings do not systematically bias
the model fits because they occur randomly over the
transit chords. Figure 6 shows 100 representative draws
from the chains that illustrate the range of allowed tran-
sit times. The dominant TTV pattern is sinusoidal with
P (cid:48)
≈ 800 days, but other harmonics are visible. To fa-
cilitate future observations of these planets we have in-
cluded our predicted transit times 2029 in the Appendix.
The planet parameters are summarized in Table 4.2.
We have included a the joint posterior distributions for
−30−20−100102030RV [m s−1]−0.4−0.20.00.20.4Phase−20−15−10−505101520RV [m s−1]b)Pb = 7.92 daysKb = 11.37 m s−1eb = 0.00 −0.4−0.20.00.20.4Phase−20−15−10−505101520c)Pc = 11.90 daysKc = 0.44 m s−1ec = 0.00 −0.4−0.20.00.20.4Phase−20−15−10−505101520d)Pd = 2.51 daysKd = -0.15 m s−1ed = 0.00 201620172018Yeara)all parameters along with a discussion of several note-
worthy covariances in the Appendix A. We found that
K2-19b and c are 32.4 ± 1.7 M⊕ and 10.8 ± 0.6 M⊕,
respectively.
While TTVs and RVs in principle provide complemen-
tary information, in our case, the TTVs are far more con-
straining. As an experiment, we ran the photodynamical
model with no RV mass priors. The mass and eccentric-
ity constraints are all consistent to within 2σ. In par-
ticular, photometry alone yields Mp,c = 30.7 ± 1.5M⊕.
We note that K2-19c is one of roughly a dozen planets
with independent mass constraints from TTVs and RVs.
See Mills & Mazeh (2017) for further discussion and a
comparison of the two techniques.
We show the constraints on the planets' eccentricity
vectors (e cos ω, e sin ω) in Figure 7. Both K2-19b and
c have moderate eccentricities of eb = 0.20 ± 0.03 and
ec = 0.21 ± 0.03 and well-aligned apsides ωc − ωb =
2 ± 2 deg. The eccentricities and orbital alignment of
these two planets have important implications for for-
mation history and their present-day dynamics, which
we discuss in Section 6.
Previously, Barros et al. (2015) measured masses
and eccentricities of Mp,b = 44 ± 12 M⊕ and Mp,c =
16.9+7.7−2.8 M⊕ and eb = 0.119+0.082
−0.035 and ec = 0.095+0.073
−0.035
using just the K2 photometry and three ground-based
transits of K2-19b. Our measurements are consistent
with those of Barros et al. (2015) at the 1 -- 2σ level, but
our measurements have smaller uncertainties on all pa-
rameters due to the additional Spitzer transits.
5. CORE/ENVELOPE STRUCTURE
Here, we examine the K2-19 planets in the context of
other known exoplanets. Figure 9 shows a mass-radius
diagram constructed from the NASA Exoplanet Archive
(Akeson et al. 2013). Our ∼5% mass measurements are
among the most precise for any sub-Jovian size planet.
Mass and radius reflect a planet's core/envelope distri-
bution. K2-19 are both "sub-Saturns," which we de-
fine as planets with Rp = 4 -- 8 R⊕. The bulk composi-
tion of sub-Saturns may be well-approximated by a two-
component model consisting of a high density core and
a H/He envelope of solar composition (Lopez & Fortney
2014; Petigura et al. 2016). For sub-Saturns, their to-
tal size is determined largely by their envelope fraction
fenv = Menv/Mp, and thus changes in the detailed core
composition weakly affect the total size.
Lopez & Fortney (2014) computed planet radii over
a grid of Mp, fenv, age, and incident flux Sinc. As a
point of reference, we show the mass-radius relationship
Table 3. K2-19 System Parameters
9
Value Notes
5322 ± 100 A
4.51 ± 0.08 A
0.06 ± 0.05 A
11.2 ± 0.03 B
3.42 ± 0.06 C
0.88 ± 0.03 D,E
0.82 ± 0.03 D,E
Parameter
Stellar Parameters
Teff (K)
log g (dex)
[Fe/H] (dex)
K (mag)
π(cid:63) (mas)
Photodynamical Analysis
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
q1
q2
Pb (days)
Tc,b (BJD−2454833)
√
√
eb cos ωb
eb sin ωb
ib (deg)
Ωb (deg)
Rp,b/R(cid:63)
Mp,b (M⊕)
Pc (days)
Tc,c (BJD−2454833)
√
√
ec cos ωc
ec sin ωc
ic (deg)
Ωc (deg)
Rp,c/R(cid:63)
Mp,c (M⊕)
Pd (days)
Tc,d (BJD−2454833)
√
√
ed cos ωd
ed sin ωd
id (deg)
Ωd (deg)
Rp,d/R(cid:63)
Mp,d (M⊕)
Derived Parameters
Rp,b (R⊕)
Rp,d (R⊕)
Rp,c (R⊕)
eb
ec
∆ω (deg)
fenv,b (%)
fenv,c (%)
Mcore,b (M⊕)
Mcore,c (M⊕)
0.4 ± 0.1 D
0.3 ± 0.2 D
7.9222 ± 0.0001 D
2027.9023 ± 0.0002 D
0.02 ± 0.06 D
−0.44 ± 0.04 D
91.5 ± 0.1 D
0 (fixed) D
0.0777 ± 0.0006 D
32.4 ± 1.7 D,F
11.8993 ± 0.0008 D
2020.0007 ± 0.0004 D
0.04 ± 0.04 D
−0.46 ± 0.03 D
91.1 ± 0.1 D
−7.4 ± 0.8 D
0.0458 ± 0.0004 D
10.8 ± 0.6 D,F
2.5081 ± 0.0002 D
2021.0726 ± 0.0018 D
0 (fixed) D
0 (fixed) D
90.8 ± 0.7 D
0 (fixed) D
0.0124 ± 0.0004 D
<10 D,F
7.0 ± 0.2 G
1.11 ± 0.05 G
4.1 ± 0.2 G
0.20 ± 0.03 G
0.21 ± 0.03 G
2 ± 2 G
44 ± 3 H
14 ± 1 H
18 ± 1 H
9.4 ± 0.5 H
Note -- A: Brewer et al. (2016). B: 2MASS (Skrutskie et al. 2006).
C: Gaia DR2 (Gaia Collaboration et al. 2018). D: Input parame-
ters into photodynamical model, see Section 4.2. E: We imposed
Gaussian priors on M(cid:63) and R(cid:63) using the methodology described
in Fulton & Petigura (2018) that incorporated A, B, and C. F:
Based on our RV analysis (Section 4.1), we imposed the following
Gaussian priors on planet masses: Mp,b = 33 ± 5 M⊕, Mp,c =
0 ± 10 M⊕, Mp,d = 0.0 ± 3.2 M⊕. G: Derived from the posterior
samples of D. H: Derived from the planet mass and radius con-
straints along with the core-envelope models of Lopez & Fortney
(2014). See Section 5 for further details.
10
Figure 6. Top: black points show measured transit times with respect to a reference linear ephemeris. Blue and orange lines
show transit times of K2-19b and c, respectively, computed from 100 draws from our MCMC chains (see Section 4.2). We do
not show times from the K2 epoch (t = 1980 -- 2060 days) because we model the flux timeseries directly. The lines in the bottom
panels represent the residuals to the predicted transit times and the formal timing uncertainties. Most of the model draws are
within 2σ of the measured transit times and indicate good agreement between data and model.
for these models at several values of fenv in Figure 8.2
Both K2-19b and c require volumetrically significant en-
velopes to explain their masses and sizes. Following Pe-
tigura et al. (2017), we derived core masses and enve-
lope fractions for these planets by interpolating over the
Lopez & Fortney (2014) model grid. K2-19c has a core
mass of 9.4 ± 0.5 M⊕ and is 14 ± 1% envelope by mass,
while K2-19b has a core mass of 18±1 M⊕ and is 44±3%
envelope by mass.
Petigura et al. (2017) compiled a sample of 23 sub-
Saturns with well-measured masses and radii to examine
trends within this population. One trend is that sub-
Saturns have a range of envelope fractions, and that
range broadens with decreasing equilibrium tempera-
ture. This broadening is likely due to the decreasing
importance of photoevaporation at lower Teq. The K2-
19 planets have intermediate Teq of ∼800 K and span
the full range of fenv.
Petigura et al. (2017) also noted a positive correlation
between the host star metallicity and the total mass of
sub-Saturns. As intermediate mass sub-Saturns around
a near solar-metallicity star, the K2-19 planets also con-
form to this trend. The emerging Mp -- [Fe/H] correlation
may point to metallicity dependent effects in the growth
of cores and/or accretion of gas from the protoplanetary
disk. However, an expanded sample size is needed to
more thoroughly assess the significance of this correla-
tion and possible dependencies on quantities like stellar
mass, which is covariant with metallicity.
With fenv = 44 ± 3%, K2-19b is one of the most
envelope-rich sub-Saturns known. Its envelope fraction
is nearly as high as K2-24c with fenv = 52+5−3% (Petigura
et al. 2018a). Like K2-24c, K2-19b presents an intrigu-
ing challenge to traditional core-accretion theory. As a
point of reference, in the canonical core accretion models
of Pollack et al. (1996), Saturn forms first as a ≈12 M⊕
core that accretes H/He from the protoplanetary disk.
2 Formally, we set age = 5 Gyr and Sinc = 80 S⊕ in order to plot
single lines, but we note that these curves are nearly overlapping
at low Sinc and late times.
−10−50510TTV(hours)−0.10.00.12000250030003500400045005000BJD-2545833−0.10.00.1Residuals(hours)11
We simulated the plausible long-term evolution of K2-
19b and c by taking 100 draws from the posterior sam-
ples from Section 4.2 and evolving them for 50 years
using the IAS15 N-body integrator included in the RE-
BOUND package (Rein & Liu 2012; Rein & Spiegel
2015).
Our integrations all revealed the same qualitative ap-
sidal outcome: circulation rather than libration of (cid:104)φ(cid:105).
In Figure 10, we show the evolution of φ for a represen-
tative simulation. The quantity 3nc − 2nb has a time
average of −0.003 rad/day, much larger than b or
c.
Instead, the planet eccentricities evolve secularly over
a period of roughly six years while the apsides remain
aligned.
In our simulations we did not include precession from
general relativity or the quadrupole field due to K2-19d.
Here, we justify these approximations. Planet b experi-
ences apsidal precession due to an effective quadrupole
moment from planet d. The rate of this precession is
given by
(cid:18) ad
(cid:19)2
ab
ωJ2 = 3nbJ2
where
We find that
J2 =
1
2
md
M(cid:63)
.
τJ2 = 2π/ ωJ2 ≈ 2 × 104 yr.
K2-19b also experiences apsidal precision due to GR
with a rate of
ωGR = 3nb
GM(cid:63)
abc2
so that
τGR = 2π/ ωGR ≈ 6 × 104 yr.
Because τGR and τJ2 are much longer than the secular
eccentricity oscillations, we are justified in neglecting
their effects above.
7. FORMATION
An intriguing aspect of the K2-19 system is that both
the physical and orbital characteristics of planets b and
c are peculiar, especially when viewed against the back-
drop of other well-characterized planetary systems, in-
cluding our own. In particular, from the perspective of
conventional planet formation theory (Armitage 2010),
the inferred properties of the K2-19 planets present a
formidable challenge. As already mentioned above, the
near-unity envelope-to-core mass fraction of K2-19c is
not a natural outcome of core-nucelated accretion model
of planet formation (Pollack et al. 1996; Hubickyj et al.
Figure 7. 2D joint posterior of e cos ω and e sin ω for K2-19b
(blue) and K2-19c (green); the contours show 1 and 2 sigma
levels. The eccentricities of both planets (i.e.
the radial
distance from the origin) are non-zero and are consistent to
within errors, eb = 0.20 ± 0.03 and ec = 0.21 ± 0.03. The
planets have well-aligned apsides with ∆ω = 2 ± 2 (deg).
At the crossover mass (i.e. when Menv ≈ Mcore or
when fenv ≈ 50%), runaway accretion begins and Saturn
quickly grows to its final mass.
One could attempt to resolve the fenv ≈ 50% prob-
lem by imagining that the disk dissipated right as K2-
19b approached the runaway phase. While this scenario
is impossible to rule out, it requires special timing of
planet formation and is thus a priori unlikely. More
likely, the inferred structure of K2-19b points to an in-
complete understanding of core-nucleated accretion and
motivates further theoretical explanations of planet con-
glomeration in the sub-Saturn mass regime.
6. MEAN-MOTION RESONANCE
K2-19b and c are clearly near the 3:2 mean-motion
resonance, but are they actually in resonance? Reso-
nance requires the libration of a resonant angle, e.g.,
φ = 3λc − 2λb − ,
where λ is the mean longitude and is the longitude of
periastron for either planet b or c. Librating angles are
confined to a particular range while circulating angles
sweep out all values between 0 and 2π. If φ is librating,
(cid:104) φ(cid:105) = 3nc − 2nb − = 0.
−0.30−0.150.000.150.30ecosω−0.30−0.150.000.150.30esinωK2-19bK2-19c12
2005). However, even if we ignore the physical struc-
ture of these planets altogether, their orbital architec-
ture lies in sharp contrast with with theoretical expec-
tations (Kley & Nelson 2012).
The most noteworthy feature of the K2-19bc pair is
their proximity to exact 3:2 mean motion commensu-
rability. In general, orbital resonances have long been
recognized as an aftereffect of convergent orbital migra-
tion (Tanaka et al. 2002; Bitsch et al. 2015). Further-
more, theoretical treatment of migration predicts that
planets as massive as K2-19b and c should have read-
ily experienced disk-driven orbital decay. Therefore, it
is not unreasonable to anticipate a distinctly resonant
present-day architecture of K2-19 that could in turn
be attributed to a migratory origin. Moreover, cou-
pled with long-range migration, resonant interactions
are well-known to adiabatically excite the orbital ec-
centricities of the constituent planets (see, e.g., Burns
& Matthews 1986; Malhotra 1995; Lee & Peale 2002),
and our photodynamical model revealed significant ec-
centricities of e ≈ 0.2. Nevertheless, as we showed in
Section 6, the system is incompatible with mean-motion
resonance, and thus the entire aforementioned narrative.
Both the values of the eccentricities themselves, as well
as the apsidal orientations of the orbits are contradictory
to those that would have been sculpted by convergent
migration. More specifically, within the framework of
the standard resonance capture scenario, orbital eccen-
tricities are determined by a balance between adiabatic
excitation that arises from convergent orbital evolution
and disk-driven eccentricity damping. Quantitatively,
this balance yields eccentricities of e ∼ h/r ∼ 0.05,
where h is the disk scale height and r is the distance
to the host star (Pichierri et al. 2018).
However, the inferred eccentricities of K2-19b and c
exceed this characteristic value by a factor of a few.
More dramatically, a clear consequence of adiabatic res-
onance capture is the anti-alignment of planetary apsi-
dal lines, such that ∆ ≈ 180 deg (Batygin & Morbidelli
2013a). Instead, in this system, the data clearly points
to apsidal alignment, characterized by ∆ ≈ 0 deg. It
is this requirement for the periapse alignment that pre-
vents us from finding a suitable resonant solution for the
planetary orbits.
To elaborate on apsidal alignment further, we note
that stable resonant equilibria that exist far away from
∆ ≈ 180 deg are indeed possible at sufficiently high
eccentricities (Beaugé et al. 2006). In an effort to con-
sider this possibility for K2-19, we carried out an N-
body numerical experiment, simulating the convergent
migration and subsequent resonant locking of K2-19b
and c. In particular, we initialized both planets on cir-
cular orbits, at a period ratio 20% outside of nominal
3:2 commensurability and computed the orbital evolu-
tion resulting from mutual gravitational perturbations
as well as a fictitious force designed to mimic planet-
disk interactions. The integration was carried out using
the Bulirsch-Stoer algorithm (Press et al. 1992), with an
accuracy parameter set to 10−10.
We adopted the model disk acceleration formulae
spelled out in Papaloizou & Larwood (2000), setting the
convergent migration timescale τa = 2 × 104 yr. While
our choice of τa was arbitrary, the resulting evolution is
adiabatic and thus insensitive to the adopted τa (Hen-
rard 1982). To prevent the system from equilibrating
in resonance with low eccentricities (e.g. Pichierri et al.
2017), we unphysically set the timescale for eccentricity
damping to τe = ∞, such that disk-driven convergent
migration resulted in continued adiabatic enhancement
of the planetary eccentricities once a resonant coupling
was established (Lee 2004).
The initial results of our simulations followed a fa-
miliar pattern: the planets migrated convergently, were
captured into the 3:2 mean-motion resonance, and de-
veloped finite eccentricities while locked into strict ap-
sidal anti-alignment with ∆ = 180◦. Once the plane-
tary eccentricities reached sufficiently large values, how-
ever, we observed deviations away from exact apsidal
anti-alignment. Nevertheless, we found that in order to
attain ∆ even remotely close to zero, unreasonably
large eccentricities were required. For example, a reso-
nant equilibrium at ∆ ∼ 60◦ requires e > 0.8 for both
planets. Thus, our results show that although asymmet-
ric equilibria can follow after capture into mean-motion
resonance, the required eccentricities are simply too high
to be observationally permissible. Indeed, resonant cou-
pling appears to be strictly ruled out by the available
data.3
For completeness, we can also speculate regarding an
alternative mechanism for finite eccentricity excitation:
mean-motion resonance crossing due to divergent mi-
gration.
In this scenario, planets start out interior to
a resonant period ratio and cross a commensurability,
which results in a non-capturing encounter with the res-
onant separatrix. This yields an impulsive excitation of
the planetary eccentricities. For example, models of the
early solar system by Tsiganis et al. (2005), planetesi-
3 As a corollary, we note that orbits which originate in resonance
can be driven out of exact commensurability while maintaining
libration of resonant angles by long-term energy dissipation (Lith-
wick & Wu 2012; Batygin & Morbidelli 2013b). This scenario,
however is only relevant to systems with vanishingly low eccen-
tricities and period ratios well outside of the nominal resonance
width, both of which are not satisfied in K2-19.
mal scattering by Jupiter and Saturn leads to divergent
migration, and the crossing of the 2:1 resonance excites
eccentricities of ≈5 -- 10%. This scenario, however, also
yields strict apsidal anti-alignment after the encounter
(Batygin & Morbidelli 2013a) and is therefore also ruled
out by the observations.
We conclude this section with a brief remark on dy-
namical stability and its relationship to the observed
orbital architecture of the K2-19 system. A trivial
examination of the derived orbital elements illustrates
that this systems is strongly AMD-unstable (Petit et al.
2018). So how is the stability of these planets ensured?
It is well known that highly eccentric planets or satellites
locked into orbital commensurabilities often derive long-
term orbital stability from the resonant phase-protection
mechanism. As we have demonstrated above, however,
in the case of K2-19b and c, libration of resonant angles
appears to be forbidden by the observational data. In-
stead, the planets around K2-19 appear to be protected
from close encounters primarily by the fact that the or-
bits have persistently co-linear apses and are therefore
geometrically nested. While this configuration is indeed
long-term stable, the dynamical genesis of this orbital
configuration remains elusive.
8. CONCLUSIONS
The K2-19 system offers a sharp contrast to the ar-
chitecture and physical properties of the solar system
planets. In the solar system, not a single planet resides
interior to Mercury (P = 88 d), while for K2-19 there
are (at least) three planets with P < 12 d. K2-19c
straddles a gap in the size distribution of solar system
planets between the ice giants and Jovians. Finally, no
pair of major solar system planets resides so close to
mean-motion resonance, although numerous Kuiper belt
objects are in resonance with Neptune, of which Pluto
is the prototypical example.
The planets orbiting K2-19 are also unusual compared
to typical extrasolar planets. Highly irradiated planets
between size of Neptune and Saturn are rare: Petigura
et al. (2018b) performed a demographic analysis of GK
stars observed by Kepler and found 0.36 planets per
100 stars with Rp = 4 -- 8 R⊕ and P < 10 d. In addition,
such proximity to resonance is not a common feature
of extrasolar planets; to first order, planet period ratios
are uniformly distributed (Lissauer et al. 2011).
Motivated by the unusual characteristics of the K2-19
planets, our team collected RVs with Keck/HIRES and
additional photometry from Spitzer and LCO. The RV
dataset was sufficient to detect the reflex motion due to
13
Figure 8. The K2-19 planets viewed in context with other
exoplanets. Gray points show the masses and radii of ex-
oplanets where mass is measured to 25% or better. The
K2-19 planets are shown in red and the uncertainties are
comparable to the point size. The blue lines show mass-
radius relationships for model planets having an Earth-
composition core and various envelope fractions of H/He,
fenv = Menv/Mp.
K2-19b at 7σ. However, the RVs alone were insufficient
to detect K2-19c due to its lower mass. Quasiperiodic
RV variability due to spots of ≈7 m s−1 limited the
sensitivity the RV dataset. Spot contrasts are smaller
at redder wavelengths, and K2-19 would benefit from
RV monitoring in the NIR by instruments such as IRD
(Kotani et al. 2018).
The high precision of the K2 and Spitzer photom-
etry combined with our multi-year time baseline pro-
vided much more stringent constraints on the physical
and orbital properties of the planets. We measured the
masses of both K2-19b and c to ≈5%, which are among
the most precise of any sub-Jovian exoplanet. Our mass
and radius measurements provided a window into the
core-envelope structure of these planets. We found that
K2-19c is roughly 15% envelope by mass, while K2-19b is
nearly 50% -- close to the canonical cross-over mass lead-
ing to runaway accretion (Pollack et al. 1996). These
planets contribute to an emerging picture of planets be-
tween size of Neptune and Saturn: where cores of a given
mass exhibit a wide diversity of envelope fractions and
where that diversity grows with decreasing irradiation
(see Figure 9).
1310301003001000Planet mass (Earth-Masses)100101Planet size (Earth-Radii)K2-19b K2-19c fenv = 10%fenv = 20%fenv = 50%14
Figure 9. The K2-19 planets viewed in context amongst other sub-Saturns (Rp = 4 -- 8 R⊕). Left: Envelope fraction vs.
equilibrium temperature. Gray points are drawn from Petigura et al. (2017). The K2-19 planets straddle the range of observed
envelope fractions. K2-19b resides near the upper envelope of the fenv distribution, which broadens toward lower Teq. Right:
Same sample as left, but showing planet mass vs. host star metallicity. There is a positive correlation between [Fe/H] and Mp,
with significant scatter. The K2-19 planets conform to this trend.
Figure 10. The dynamical evolution of a representative solution from our photodynamical model, based on a 50 year N-body
integration. Left: Eccentricity as a function of time. The planets exchange eccentricity over a secular timescale of ∼6 years.
Middle: same simulation as left, but with ∆ on the x-axis. The planets precess together and retain apsidal alignment. Right:
Several angular velocities relevant to the planet's resonant state. Because the quantity 3nc − 2nb − does not have a time
average of zero for either b or c, the planets are not in the 3:2 mean motion resonance. Instead, the resonant angles circulate
at a rate of ∼1 radian per year.
Through our photodynamical analysis, we found that
these planets have moderate eccentricities of ≈0.2 and
aligned apsides. The planets are experiencing rapid sec-
ular eccentricity oscillations with a ≈6 yr timecale, but
the system is currently not in mean-motion resonance.
Moreover, the system's present configuration presents
a challenge to formation pathways that involve mean-
motion resonance in the past. Scenarios where the sys-
tem passes through the 3:2 resonance from above or be-
low predict anti-aligned apsides, which are ruled out by
the data. Future photometric or RV monitoring would
shed additional light on this enigmatic system.
50010001500Equilibrium Temp (K)020406080100Envelope Fraction (%)K2-19bK2-19cK2-24 c−0.20.00.20.4[Fe/H] (dex)131030100Planet Mass (Earth-masses)K2-19b K2-19c 02000040000BJD - 24548330.000.050.100.150.200.250.30EccentricityK2-19bK2-19c−1000100Δϖ (deg)0.000.050.100.150.200.250.30EccentricityK2-19bK2-19c02000040000BJD - 2454833−0.0100−0.0075−0.0050−0.00250.00000.00250.00500.00750.0100Angular Velocity (rad/day)(3nc−2nb)ϖbϖcACKNOWLEDGMENTS
We thank the anonymous reviewer for helpful sugges-
tions that improved this manuscript.
for Research in Astronomy,
E.A.P. acknowledges support for this work by NASA
through the NASA Hubble Fellowship grant HST-
HF2-51417 awarded by the Space Telescope Science
Institute, which is operated by the Association of
Universities
for
NASA, under contract NAS5-26555.
S.C.C.B. ac-
knowledges support from Fundação para a Ciência e
a Tecnologia (FCT) through Investigador FCT con-
tract IF/01312/2014/CP1215/CT0004 and by FEDER
through COMPETE2020 and POCI in the framework of
the project POCI-01-0145-FEDER-028953.
Inc.,
This work made use of NASA's Astrophysics Data
System Bibliographic Services.
This work is based in part on observations made with
the Spitzer Space Telescope, which is operated by the
Jet Propulsion Laboratory, California Institute of Tech-
15
nology under a contract with NASA. Support for this
work was provided by NASA through an award issued
by JPL/Caltech. This work makes use of observations
from the LCOGT network.
Some of the data presented herein were obtained at the
W. M. Keck Observatory, which is operated as a scien-
tific partnership among the California Institute of Tech-
nology, the University of California, and NASA. The
authors wish to recognize and acknowledge the very sig-
nificant cultural role and reverence that the summit of
Maunakea has always had within the indigenous Hawai-
ian community. We are most fortunate to have the op-
portunity to conduct observations from this mountain.
(HIRES),
Spitzer, Keck:I
Facilities:
Kepler,
LCOGT
Software:
batman
emcee
(Foreman-Mackey et al. 2013), Phodymm (Mills et al.
2016), REBOUND (Rein & Liu 2012), EVEREST2.0
(Luger et al. 2017), RadVel (Fulton et al. 2018).
(Kreidberg 2015),
REFERENCES
Aigrain, S., Pont, F., & Zucker, S. 2012, MNRAS, 419, 3147
Akeson, R. L., Chen, X., Ciardi, D., et al. 2013, PASP, 125,
989
Armitage, P. J. 2010, Astrophysics of Planet Formation
Armstrong, D. J., Santerne, A., Veras, D., et al. 2015,
ArXiv e-prints, arXiv:1503.00692
Deck, K. M., Agol, E., Holman, M. J., & Nesvorný, D.
2014, ApJ, 787, 132
Deming, D., Knutson, H., Kammer, J., et al. 2015, ApJ,
805, 132
Eastman, J., Gaudi, B. S., & Agol, E. 2013, PASP, 125, 83
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman,
Barros, S. C. C., Almenara, J. M., Demangeon, O., et al.
J. 2013, PASP, 125, 306
2015, MNRAS, 454, 4267
Fulton, B. J., & Petigura, E. A. 2018, ArXiv e-prints,
Batygin, K., & Morbidelli, A. 2013a, A&A, 556, A28
-- . 2013b, AJ, 145, 1
Beaugé, C., Michtchenko, T. A., & Ferraz-Mello, S. 2006,
MNRAS, 365, 1160
arXiv:1805.01453
Fulton, B. J., Petigura, E. A., Blunt, S., & Sinukoff, E.
2018, PASP, 130, 044504
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al.
Benneke, B., Werner, M., Petigura, E., et al. 2017, ApJ,
2018, A&A, 616, A1
834, 187
Bitsch, B., Lambrechts, M., & Johansen, A. 2015, A&A,
582, A112
Brewer, J. M., Fischer, D. A., Valenti, J. A., & Piskunov,
N. 2016, ApJS, 225, 32
Gelman, A., & Rubin, D. B. 1992, Statistical Science, 7, 457
Goodman, J., & Weare, J. 2010, Communications in
Applied Mathematics and Computational Science, 5, 65
Grunblatt, S. K., Howard, A. W., & Haywood, R. D. 2015,
ApJ, 808, 127
Brown, T. M., Baliber, N., Bianco, F. B., et al. 2013,
Haywood, R. D., Collier Cameron, A., Queloz, D., et al.
PASP, 125, 1031
Burns, J. A., & Matthews, M. S., eds. 1986, Orbital
resonances, unusual configurations and exotic rotation
statesamong planetary satellites., ed. J. A. Burns &
M. S. Matthews, 159 -- 223
2014, MNRAS, 443, 2517
Henrard, J. 1982, Celestial Mechanics, 27, 3
Howard, A. W., & Fulton, B. J. 2016, PASP, 128, 114401
Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010,
ApJ, 721, 1467
Claret, A., Hauschildt, P. H., & Witte, S. 2012, A&A, 546,
Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126,
A14
398
Dai, F., Winn, J. N., Albrecht, S., et al. 2016, ApJ, 823, 115
Deck, K. M., & Agol, E. 2015, ApJ, 802, 116
Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2005,
Icarus, 179, 415
16
Ingalls, J. G., Krick, J. E., Carey, S. J., et al. 2012, in
Petigura, E. A., Sinukoff, E., Lopez, E. D., et al. 2017, AJ,
Proc. SPIE, Vol. 8442, Space Telescopes and
Instrumentation 2012: Optical, Infrared, and Millimeter
Wave, 84421Y
Kipping, D. M. 2013, MNRAS, 435, 2152
Kley, W., & Nelson, R. P. 2012, ARA&A, 50, 211
Kotani, T., Tamura, M., Nishikawa, J., et al. 2018, in
Society of Photo-Optical Instrumentation Engineers
(SPIE) Conference Series, Vol. 10702, Proc. SPIE,
1070211
Kreidberg, L. 2015, PASP, 127, 1161
Lee, M. H. 2004, ApJ, 611, 517
Lee, M. H., & Peale, S. J. 2002, ApJ, 567, 596
Lissauer, J. J., Ragozzine, D., Fabrycky, D. C., et al. 2011,
ApJS, 197, 8
Lithwick, Y., & Wu, Y. 2012, ApJL, 756, L11
Lithwick, Y., Xie, J., & Wu, Y. 2012, ApJ, 761, 122
Livingston, J. H., Crossfield, I. J. M., Werner, M. W., et al.
2019, AJ, 157, 102
Lomb, N. R. 1976, Ap&SS, 39, 447
Lopez, E. D., & Fortney, J. J. 2014, ApJ, 792, 1
Luger, R., Kruse, E., Foreman-Mackey, D., Agol, E., &
Saunders, N. 2017, ArXiv e-prints, arXiv:1702.05488
Malhotra, R. 1995, AJ, 110, 420
Marcy, G. W., & Butler, R. P. 1992, PASP, 104, 270
McCully, C., Volgenau, N. H., Harbeck, D.-R., et al. 2018,
in Software and Cyberinfrastructure for Astronomy V,
Vol. 10707, 107070K
Mills, S. M., Fabrycky, D. C., Migaszewski, C., et al. 2016,
Nature, 533, 509
Mills, S. M., & Mazeh, T. 2017, ApJL, 839, L8
Narita, N., Hirano, T., Fukui, A., et al. 2015, ApJ, 815, 47
Nespral, D., Gandolfi, D., Deeg, H. J., et al. 2017, A&A,
601, A128
Pál, A. 2008, MNRAS, 390, 281
Papaloizou, J. C. B., & Larwood, J. D. 2000, MNRAS, 315,
823
Petigura, E. A., Howard, A. W., Lopez, E. D., et al. 2016,
ApJ, 818, 36
153, 142
Petigura, E. A., Benneke, B., Batygin, K., et al. 2018a, AJ,
156, 89
Petigura, E. A., Marcy, G. W., Winn, J. N., et al. 2018b,
AJ, 155, 89
Petit, A. C., Laskar, J., & Boué, G. 2018, A&A, 617, A93
Pichierri, G., Morbidelli, A., & Crida, A. 2018, Celestial
Mechanics and Dynamical Astronomy, 130, 54
Pichierri, G., Morbidelli, A., & Lai, D. 2017, A&A, 605, A23
Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996,
Icarus, 124, 62
Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231
Press, W. H., Teukolsky, S. A., Vetterling, W. T., &
Flannery, B. P. 1992, Numerical recipes in FORTRAN.
The art of scientific computing
Rein, H., & Liu, S.-F. 2012, A&A, 537, A128
Rein, H., & Spiegel, D. S. 2015, MNRAS, 446, 1424
Scargle, J. D. 1982, ApJ, 263, 835
Schwarz, G. 1978, Annals of Statistics, 6, 461
Sinukoff, E., Howard, A. W., Petigura, E. A., et al. 2016,
ApJ, 827, 78
Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ,
131, 1163
Tanaka, H., Takeuchi, T., & Ward, W. R. 2002, ApJ, 565,
1257
Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F.
2005, Nature, 435, 459
Valenti, J. A., Butler, R. P., & Marcy, G. W. 1995, PASP,
107, 966
VanderPlas, J. T. 2018, ApJS, 236, 16
Vaughan, A. H., Preston, G. W., & Wilson, O. C. 1978,
PASP, 90, 267
Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, 2198,
362
Winn, J. N., Holman, M. J., Torres, G., et al. 2008, ApJ,
683, 1076
17
APPENDIX
A. PHOTODYNAMICAL MODEL
Here, we include some supplemental information regarding our photodynamical model described in Section 4.2.
Table 4 lists the predicted transit times and uncertainties for K2-19b and c up to 2029. Figure 11 shows the 2D joint
posteriors of all parameters included in our photodynamical model.
We highlight the covariances between mass and eccentricity in Figure 12. The masses of planet b and c are correlated
because the amplitudes of near-resonant TTVs constrain planet mass ratios (Lithwick et al. 2012). However, the RVs
and higher order TTV terms (i.e. chopping) constrain the individual masses directly (Deck & Agol 2015). Figure 12
also illustrates a positive correlation between eb cos ωb and ec cos ωc and between eb sin ωb and ec sin ωc. This is another
common feature of near-resonant systems: the TTV amplitude and phase encodes linear combinations of e cos ω and
e sin ω (Lithwick et al. 2012).
Table 4. Predicted Transit Times
Planet
i
UTC date
b
c
b
c
b
c
b
b
c
b
0
0
1
1
2
476
716
717
477
718
2014-06-04
2014-06-08
2014-06-12
2014-06-20
2014-06-20
2029-12-11
2029-12-14
2029-12-21
2029-12-23
2029-12-29
Tc
days
1980.3840
1984.2722
1988.3041
1996.1834
1996.2220
7648.8365
7651.5243
7659.4466
7660.7298
7667.3662
σ(Tc)
days
0.0002
0.0008
0.0002
0.0006
0.0002
0.1814
0.0468
0.0446
0.1710
0.0434
Note -- Predicted transit times for K2-19b and c,
where i, is an index that labels individual transits.
Times are given in BJDTBD − 2454833. Table 1
is published in its entirety in the machine-readable
format. A portion is shown here for guidance re-
garding its form and content.
18
Figure 11. 2D joint posterior probability distributions for our photodynamical model (Section 4.2). The dark and light regions
show 1 and 2 sigma contours, respectively.
2021.0682021.0712021.0742021.077Tc,d(days)90.491.292.092.893.6id(deg)36912Mp,d(M⊕)0.01140.01200.01260.0132Rp,d/R?7.92187.92207.92227.9224Pp,b(days)2027.90172027.90202027.90232027.90262027.9029Tc,b(days)−0.150.000.15√ebcosωb−0.56−0.48−0.40−0.32−0.24√ebsinωb91.291.491.691.8ib(deg)2730333639Mp,b(M⊕)0.0760.0770.0780.079Rp,b/R?11.897511.899011.900511.9020Pp,c(days)2020.00002020.00052020.00102020.00152020.0020Tc,c(days)−0.16−0.080.000.080.16√eccosωc−0.54−0.48−0.42−0.36−0.30√ecsinωc−9.0−7.5−6.0−4.5Ωc(deg)910111213Mp,c(M⊕)0.04440.04500.04560.04620.0468Rp,c/R?0.750.800.850.90R?0.800.850.900.95M?0.150.300.45q12.50762.50802.50842.5088Pp,d(days)0.20.40.60.8q22021.0682021.0712021.0742021.077Tc,d(days)90.491.292.092.893.6id(deg)36912Mp,d(M⊕)0.01140.01200.01260.0132Rp,d/R?7.92187.92207.92227.9224Pp,b(days)2027.90172027.90202027.90232027.90262027.9029Tc,b(days)−0.150.000.15√ebcosωb−0.56−0.48−0.40−0.32−0.24√ebsinωb91.291.491.691.8ib(deg)2730333639Mp,b(M⊕)0.0760.0770.0780.079Rp,b/R?11.897511.899011.900511.9020Pp,c(days)2020.00002020.00052020.00102020.00152020.0020Tc,c(days)−0.16−0.080.000.080.16√eccosωc−0.54−0.48−0.42−0.36−0.30√ecsinωc−9.0−7.5−6.0−4.5Ωc(deg)910111213Mp,c(M⊕)0.04440.04500.04560.04620.0468Rp,c/R?0.750.800.850.90R?0.800.850.900.95M?0.150.300.45q119
Figure 12. Same as Figure 11, but highlighting several noteworthy covariances between planet masses and planet eccentricities.
Left panel: Constraint on Mp,b and Mp,c. The covariance between the planet masses is typical of TTV analyses which tend to
provide smaller fractional uncertainties on mass ratios than on individual masses. Middle panel: same as left but for eb cos ωb
and ec cos ωc. The covariance results from the fact that TTVs constrain linear combinations of the eccentricity vectors. Right
panel: same as middle but for eb sin ωb and ec sin ωc.
010203040Mp,b(M⊕)010203040Mp,c(M⊕)−0.30−0.150.000.150.30ebcosωb−0.30−0.150.000.150.30eccosωc−0.30−0.150.000.150.30ebsinωb−0.30−0.150.000.150.30ecsinωc |
1104.2028 | 1 | 1104 | 2011-04-11T19:28:03 | A Search for Satellite around Ceres | [
"astro-ph.EP"
] | We conducted a satellite search around the dwarf planet 1 Ceres using Hubble Space Telescope and ground-based Palomar data. No candidate objects were found orbiting Ceres in its entire stability region down to ~500km from the surface of Ceres. Assuming a satellite would have the same albedo as Ceres, which has a visual geometric albedo of 0.07-0.10, our detection limit is sensitive to satellites larger than 1-2 km in diameter. | astro-ph.EP | astro-ph | A Search for Satellites around Ceres
A. Bieryla1,2, J. Wm. Parker2, E.F. Young2, L. A. McFadden3, C. T. Russell4, S. A. Stern2,
M. V. Sykes5 and B. Gladman7
1Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, MS-6, Cambridge, MA
02138, USA [email protected]
2Southwest Research Institute, 1050 Walnut Street, Suite 300, Boulder, CO 80302, USA
3 Goddard Spaceflight Center, Greenbelt, MD 20771, USA
4 Institute of Geophysics and Planetary Physics, University of California, Los Angeles,
CA 90095, USA
5 Planetary Science Institute, 1700 East Fort Lowell, Suite 106, Tucson, AZ, 85719 USA
6 Department of Physics and Astronomy, University of British Columbia, 6224
Agricultural Road, Vancouver, BC V6T1Z1, Canada
Abstract
Introduction
We conducted a satellite search around the dwarf planet 1 Ceres using Hubble Space
Telescope and ground-based Palomar data. No candidate objects were found orbiting
Ceres in its entire stability region down to ~500km from the surface of Ceres. Assuming
a satellite would have the same albedo as Ceres, which has a visual geometric albedo of
0.07—0.10, our detection limit is sensitive to satellites larger than 1-2 km in diameter.
1.
1 Ceres is the smallest dwarf planet and the first known and largest main-belt asteroid.
Previous observations by Hubble Space Telescope (HST) reveal Ceres to be an oblate
spheroid with axes a=487 km and b=455km, with a surface topography that is apparently
relaxed (Thomas et al. 2005). The shape and rotation suggests Ceres is differentiated
with an ice-rich mantle (Thomas et al. 2005).
Observations of satellites around minor planets are valuable because they can be used to
probe the physical properties of the components. The total mass of the system can be
determined from the relative orbit of the satellite, and the individual masses can be
determined if the orbits of both components around the barycenter can be measured.
Further, if the sizes of one or both components can be measured (either directly or from
thermal observations), then density and albedo can be determined. Such data can give
clues to the composition and interior structure properties of a body, which can be
considered in context with its taxonomy group as determined from spectroscopy, if
available. Asteroids with satellites also allow for the study of natural collisions that are
relevant to the formation and evolution of asteroids (Merline et al. 2002). As of January
2011, there were approximately 193 small bodies with known satellites. This includes 37
near-Earth asteroids, 13 Mars crossing asteroids, 73 main-belt asteroids, 4 Trojan
asteroids, and 66 trans-Neptunian objects (Johnston 2011). To date, no results from any
dedicated Ceres satellite searches have been published.
The existence, or lack thereof, of satellites around Ceres is also of great interest to
NASA’s Dawn mission, which launched on September 27, 2007 and will arrive at Ceres
in February 2015 after first rendezvousing with Vesta in 2011. The Dawn spacecraft will
orbit Ceres for several months at altitudes of 700 to 5900 km (Russell et al. 2006), which
is well inside the satellite stability region. In addition to providing additional targets for
physical studies by the Dawn instruments, satellites could affect the orbit design in order
to accommodate observations and spacecraft safety. Therefore, advance knowledge of
the environment around Ceres is critical to mission planning.
In this paper we report on our project to search for satellites around Ceres using the
Hubble Space Telescope (HST) and the Hale 5-meter telescope at Palomar Observatory.
This combination of space- and ground-based datasets allowed us to search the full extent
of the satellite stability region around Ceres and particularly to high resolution in the
inner region.
2. Observations and Analysis
We made observations of Ceres using the High Resolution Channel (HRC) of the
Advanced Camera for Surveys (ACS) on HST to study the physical properties of Ceres
(Parker et al. 2004, Thomas et al. 2005, Li et al. 2006) and to search for satellites. We
observed Ceres with the HST ACS-HRC in December 2003 and January 2004. At that
time, Ceres was 1.65 AU from the Earth, producing a scale of 30 km/pixel, and an
angular diameter of ~ 0.8 arcsec (about 31 pixels across). A total of 267 individual
images of Ceres were obtained during nine HST orbits. The first six orbits were
consecutive, covering the extent of the 9.1 hr rotation period, and the remaining three
orbits were obtained at different times in the following month to improve rotational phase
coverage and spatial resolution. Most of the exposures on the six consecutive orbits used
a 12×12 arcsec (512×512 pixel) subarray of the HRC to provide shorter readout time and
in order to fit more images into the HST memory. At the end of four of these six orbits,
“long V” (F555W filter) exposures were taken using the full 25×25 arcsec field of view
of the ACS-HRC in order to search for possible satellites of Ceres. These four
observations were made with a 15-second exposure time and a CR-SPLIT=3 to robustly
remove cosmic rays. A satellite in these images would have a typical orbital speed of less
than ~0.25 km/s (we later show that we could detect faint satellites as close as 0.4 arcsec
from Ceres), so it will move significantly less than a resolution element over the course
of a single 15-second CR-SPLIT exposure. Thus, we can combine the three sub-
exposures for each observation to remove cosmic rays without accidentally removing any
potential satellites (the observations tracked at Ceres’ apparent motion). The images were
processed using the standard HST pipeline data reduction and IRAF and IDL programs.
The HST data were able to obtain high resolution close to Ceres, but the field of view was
considerably smaller than Ceres’ Hill sphere. As discussed by Hamilton and Burns
(1991), the Hill sphere has a radius RH = (µ/3)1/3R, where R is the circular heliocentric
orbit of the primary body, and μ=mp/(m+mp) is the reduced mass-parameter. For Ceres,
RH = 2.2x105 km which translates to an apparent Hill angular radius of about 189 arcsec
during the epoch of our HST observations. Initially-circular prograde orbits remain bound
out to about half the Hill region, RH/2 = 1.1x105 km or 95 arcsec, while initially-circular
retrograde orbits are stable out to almost the full Hill sphere. The HST images cover an
area which is only the inner ~2.5% for prograde orbits and 0.6% for retrograde orbits of
the stability region.
To get complete coverage of the satellite stability region around Ceres, we also obtained
ground-based observations on 2006 July 27 with the Large Format Camera on the Hale
5m telescope at Palomar Observatory. Each CCD of the camera covers an area of roughly
720×360 arcsec (with the larger dimension in the E-W direction). The images were
binned 2x2, producing a plate scale of 0.36 arcsec/pixel. At the time, Ceres was 2.02 AU
from the Earth, resulting in a physical scale of 530 km/pixel, and angular radii of the
prograde stability region and Hill sphere of 74 arcsec and 149 arcsec, respectively. Five
one-minute exposures using the SDSS r’ filter were taken in succession over a period of
9 minutes. The motion of Ceres was small but noticeable throughout, covering about
4 arcsec between the first and last exposures. Ceres moved 0.5 arcsec during a single
exposure, which was far less than the seeing disk of FWHM~2.2 arcsec, so the motion of
any satellite would be readily detectable by eye over the series of images with little
trailing loss in each image.
We performed similar search methods and analyses on both the HST and the Palomar
images sets. We searched for objects co-moving with Ceres by blinking the normally
processed images with various levels of contrast, and then by blinking the same set of
images after performing median-filter subtraction (high-pass filter or unsharp masking) to
remove large-scale variations due to scattered light from Ceres. These filtered images
allowed us to search closer to Ceres and to a fainter detection limit. In the HST images,
there were no other objects (e.g., stars) in the images with Ceres, so the search involved
simply looking for any object that was persistent among the images taken at different
times. In the larger Palomar images, many stars were present, and the images were
manually blinked to look for any objects that moved at the same rate as Ceres relative to
the stars.
We determined the detection limits of our search by the common method of planting
“fake” objects in the images. These objects were modeled to have the same shape as the
image point spread function (PSF), and were scaled to a range of magnitudes and placed
in the images at random positions with motion consistent with Ceres. The images were
then re-examined and any detected objects were noted. In this way, we were able to
produce detection efficiency curves as a function of magnitude and distance from Ceres
for each instrument. This also provided a second pass through the data to look for real
objects that would show up as detections that weren’t in the list of fake objects.
3. Results
We did not find any satellites around Ceres. Our results for our HST detection efficiency
are shown in Figure 1 and our Palomar detection efficiency in Figure 2. The conversion
to diameter annotated along the top axis assumes a satellite with the same albedo as
Ceres, i.e., we used the observed magnitude and diameter of Ceres to normalize the
diameter scale along the top axis.
Our HST detection limits are valid to as close as ~0.4 arcsec (a projected distance of ~480
km) off the surface of Ceres. For objects in circular orbits about Ceres, 480 km is well
inside the Roche limit of Ceres. Any satellite with an orbit smaller than that could have
escaped our detection due to scattered light from Ceres. However, we also examined the
shorter-exposure F555W images that were taken at different epochs; although those
images did not go as deep (a factor of 5 less exposure time) and covered ¼ of the area as
the 15-second images, they provided some additional check for satellites that could have
been obscured by scattered light from Ceres in the 15 second images. Also, any satellite
in a low inclination orbit could have been obscured by occultations/transits with Ceres,
but we calculate that based on the timing of our observations relative to the range of
orbital periods, any satellite with an orbit larger than ~825 km would have appeared in at
least two of our images. There is a 35% chance a satellite on a smaller orbit (surface < R
< 825 km) would not have appeared in at least two of the 15-second images, but those
closer distances were obscured by scattered light anyway in the 15-second exposures.
With the Palomar data, we are not able to detect objects as close to Ceres due to lower
resolution and Ceres being over-exposed, but those ground-based images were able to
cover an area well beyond the Hill radius, providing a complementary dataset to the HST
observations covering the full stability region of potential satellites.
We are able to say (at the 90% detection level) that there are no objects larger than ~1 km
in diameter orbiting Ceres in the HST images from ~500 km above the surface out to a
distance of ~15,000 km; that size limit corresponds to a magnitude of ~21.5. With the
same detection level of 90%, we can say that there are no objects larger than ~2 km in
diameter (magnitude ~21) orbiting Ceres in the entire stability region. The two datasets
together imply that Ceres does not have any satellites larger than about 1—2 km in
diameter. These results assume a satellite would have the same albedo as Ceres; if it has
an albedo different than that of Ceres, then the diameter limit of our search would be
correspondingly different by a factor of 1/sqrt(albedo), e.g., if the albedo of a satellite
were 4 times larger than the albedo of Ceres, then our detection size limit would be 2
times smaller. Note that our results are not dependent on the specific numeric values for
the albedos of the satellite or Ceres, but only depend on the relative values of their
albedos. For reference, published values for the visual geometric albedo of Ceres are in
the range 0.07-0.10 (Li et al. 2006; Millis et al. 1987; Tedesco 1989).
The result that Ceres does not have any satellites larger than ~1 km is interesting. Unlike
Dawn’s first target Vesta, which has a large asteroid family associated with it, probably
arising from the large impact resulting in the crater covering Vesta’s southern
hemisphere, Ceres has no associated asteroid family and exhibits no large impact
structures on its surface. Ceres is alone among the asteroids in having a hydrostatic
equilibrium shape, which is maintained by its gravity and ice-rich composition. This may
offer some insight into why Ceres has no associated family or satellites: ejecta would
likely be ice-rich and because Ceres’ orbit is entirely within the snow-line around 3 AU,
they may not survive against solar radiation, much like a disintegrating cometary body.
Also, craters on an ice-rich surface would likely flatten out over time because there
would be a quicker relaxation time on a surface that is mechanically weak.
Acknowledgements. This work was supported by the HST grant from STScI (HST-GO-
09748). We would like to thank Jake Hanson for his help on the occultation analysis.
References
Hamilton, Douglas P., Burns, Joseph A. 1991. Orbital Stability Zones about Asteroids,
Icarus, 92, 118-131.
Johnston, Wm. Robert 2011. Asteroids with Satellites.
http://www.johnstonsarchive.net/astro/asteroidmoons.html
Li, J.-Y., McFadden, L.A., Parker, J.Wm., Young, E.F., Stern, S.A., Thomas, P.C.,
Russell, C.T., Sykes, M.V. 2006. Photometric Analysis of 1 Ceres and Surface
Mapping from HST observations. Icarus, 182, 143-160.
Merline, W.J., Weidenschilling, S.J., Durda, D.D., Margot, J.L., Pravec, P., and Storrs,
A.D, 2002. Asteroids Do Have Satellites? in Asteroids III, eds. W.F. Bottke, A.
Cellino, P. Paolicchi, & R.P. Binzel, Univ. of Arizona Press, 289-312
Millis, R. L., Wasserman, L. H., Franz, O. G., Nye, R. A., Oliver, R. C., Kreidl, T. J.,
Jones, S. E., Hubbard, W., Lebofsky, L., Goff, R., Marcialis, R., Sykes, M., Frecker,
J., Hunten, D., Zellner, B., Reitsema, H., Schneider, G., Dunham, E., Klavetter, J.,
Meech, K., Oswalt, T., Rafert, J., Strother, E., Smith, J., Povenmire, H., Jones, B.,
Kornbluh, D., Reed, L., Izor, K., A'Hearn, M. F., Schnurr, R., Osborn, W., Parker, D.,
Douglas, W. T., Beish, J. D., Klemola, A. R., Rios, M., Sanchez, A., Piironen, J.,
Mooney, M., Ireland, R. S., & Leibow, D. 1987, The size, shape, density, and albedo
of Ceres from its occultation of BD+8 471. Icarus, 72, 507.
Parker, J.Wm., McFadden, L.A., Russell, C.T., Stern, S.A., Sykes, M., Thomas, P.C.,
Young, E., Cline, T., Hutcheson, C. 2004. Ceres: High-resolution mapping with HST
and the determination of physical properties. 35th COSPAR Scientific Assembly,
1526.
Russell, C.T., F. Capaccioni, A. Coradini, U. Christensen, M.C. De Sanctis, W. C.
Feldman, R. Jaumann, H. U. Keller, A. Konopliv, T. B. McCord, L. A. McFadden, H.
Y. McSween, S. Mottola, G. Neukum, C. M. Pieters, T. H. Prettyman, C. A.
Raymond, D. E. Smith, M. V. Sykes, B. Williams, and M. T. Zuber, (2006) Dawn
Discovery mission to Vesta and Ceres: Present Status, Advances in Space Research
38, 2043-2048.
Tedesco, E. F. 1989, in Asteroids II, ed. R. P. Binzel, T. Gehrels, & M. S. Matthews
(Tucson: Univ. Arizona Press)
Thomas, P.C., Parker, J.Wm., McFadden, L.A., Russell, C.T., Stern, S.A., Sykes, M.V.,
Young, E.F. 2005. Differentiation of the Asteroid Ceres as Revealed by its Shape.
Nature, 437, 224-226.
|
1808.00485 | 2 | 1808 | 2018-11-22T02:05:55 | A second terrestrial planet orbiting the nearby M dwarf LHS 1140 | [
"astro-ph.EP"
] | LHS 1140 is a nearby mid-M dwarf known to host a temperate rocky super-Earth (LHS 1140 b) on a 24.737-day orbit. Based on photometric observations by MEarth and Spitzer as well as Doppler spectroscopy from HARPS, we report the discovery of an additional transiting rocky companion (LHS 1140 c) with a mass of $1.81\pm0.39~{\rm M_{Earth}}$ and a radius of $1.282\pm0.024~{\rm R_{Earth}}$ on a tighter, 3.77795-day orbit. We also obtain more precise estimates of the mass and radius of LHS 1140 b to be $6.98\pm0.89~{\rm M_{Earth}}$ and $1.727\pm0.032~{\rm R_{Earth}}$. The mean densities of planets b and c are $7.5\pm1.0~\rm{g/cm^3}$ and $4.7\pm1.1~\rm{g/cm^3}$, respectively, both consistent with the Earth's ratio of iron to magnesium silicate. The orbital eccentricities of LHS 1140 b and c are consistent with circular orbits and constrained to be below 0.06 and 0.31, respectively, with 90% confidence. Because the orbits of the two planets are co-planar and because we know from previous analyses of Kepler data that compact systems of small planets orbiting M dwarfs are commonplace, a search for more transiting planets in the LHS 1140 system could be fruitful. LHS 1140 c is one of the few known nearby terrestrial planets whose atmosphere could be studied with the upcoming James Webb Space Telescope. | astro-ph.EP | astro-ph | Draft version November 26, 2018
Typeset using LATEX preprint style in AASTeX62
8
1
0
2
v
o
N
2
2
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
5
8
4
0
0
.
8
0
8
1
:
v
i
X
r
a
A second terrestrial planet orbiting the nearby M dwarf LHS 1140
Kristo Ment,1 Jason A. Dittmann,2 Nicola Astudillo-Defru,3 David Charbonneau,1
Jonathan Irwin,1 Xavier Bonfils,4 Felipe Murgas,5 Jose-Manuel Almenara,4
Thierry Forveille,4 Eric Agol,6 Sarah Ballard,7 Zachory K. Berta-Thompson,8
Franc¸ois Bouchy,9 Ryan Cloutier,10, 11, 12 Xavier Delfosse,4 Ren´e Doyon,12
Courtney D. Dressing,13 Gilbert A. Esquerdo,1 Raphaelle D. Haywood,1
David M. Kipping,14 David W. Latham,1 Christophe Lovis,9 Elisabeth R. Newton,7
Francesco Pepe,9 Joseph E. Rodriguez,1 Nuno C. Santos,15, 16 Thiam-Guan Tan,17
Stephane Udry,9 Jennifer G. Winters,1 and Anael Wunsche4
1Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
251 Pegasi b Postdoctoral Fellow, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA
02139, USA
3Departamento de Astronom´ıa, Universidad de Concepci´on, Casilla 160-C, Concepci´on, Chile
4Universit´e Grenoble Alpes, CNRS, IPAG, F-38000 Grenoble, France
5Instituto de Astrof´ısica de Canarias (IAC), E-38200 La Laguna, Tenerife, Spain
6Astronomy Department, University of Washington, Seattle, WA 98195, USA
7Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02138, USA
8Department of Astrophysical and Planetary Sciences, University of Colorado, Boulder, CO 80309, USA
9Observatoire de l'Universit´e de Gen`eve, 51 chemin des Maillettes, 1290 Versoix, Switzerland
10Dept. of Astronomy & Astrophysics, University of Toronto, 50 St. George Street, M5S 3H4, Toronto, ON, Canada
11Centre for Planetary Sciences, Dept. of Physical & Environmental Sciences, University of Toronto Scarborough,
1265 Military Trail, M1C 1A4, Toronto, ON, Canada
12Institut de Recherche sur les Exoplan`etes, D´epartement de physique, Universit´e de Montr´eal, CP 6128 Succ.
Centre-ville, H3C 3J7, Montr´eal, QC, Canada
13Department of Astronomy, University of California, Berkeley, CA 94720, USA
14Department of Astronomy, Columbia University, 550 W 120th Street, New York, NY 10027, USA
15Instituto de Astrof´ısica e Ciencias do Espa¸co, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto,
Portugal
16Departamento de F´ısica e Astronomia, Faculdade de Ciencias, Universidade do Porto, Rua do Campo Alegre,
17Perth Exoplanet Survey Telescope, Perth, Western Australia 6010, Australia
P-4169-007 Porto, Portugal
(Received August 2, 2018; Revised September 6, 2018; Accepted November 15, 2018)
Submitted to AJ
ABSTRACT
LHS 1140 is a nearby mid-M dwarf known to host a temperate rocky super-Earth
(LHS 1140 b) on a 24.737-day orbit. Based on photometric observations by MEarth
and Spitzer as well as Doppler spectroscopy from HARPS, we report the discovery of
Corresponding author: Kristo Ment
[email protected]
2
Ment et al.
an additional transiting rocky companion (LHS 1140 c) with a mass of 1.81 ± 0.39
M⊕ and a radius of 1.282 ± 0.024 R⊕ on a tighter, 3.77795-day orbit. We also obtain
more precise estimates of the mass and radius of LHS 1140 b to be 6.98 ± 0.89 M⊕
and 1.727 ± 0.032 R⊕. The mean densities of planets b and c are 7.5 ± 1.0 g/cm3
and 4.7 ± 1.1 g/cm3, respectively, both consistent with the Earth's ratio of iron to
magnesium silicate. The orbital eccentricities of LHS 1140 b and c are consistent
with circular orbits and constrained to be below 0.06 and 0.31, respectively, with 90%
confidence. Because the orbits of the two planets are co-planar and because we know
from previous analyses of Kepler data that compact systems of small planets orbiting M
dwarfs are commonplace, a search for more transiting planets in the LHS 1140 system
could be fruitful. LHS 1140 c is one of the few known nearby terrestrial planets whose
atmosphere could be studied with the upcoming James Webb Space Telescope.
Keywords: planets and satellites: detection, terrestrial planets -- techniques: photo-
metric, radial velocities
1. INTRODUCTION
Small planets are very common around M dwarfs (Dressing & Charbonneau 2013; Bonfils et al.
2013; Mulders et al. 2015), with a cumulative occurrence rate of at least 2.5 ± 0.2 planets per M
dwarf (Dressing & Charbonneau 2015). Compared to Sun-like counterparts, such planets are easier
to detect due to the smaller sizes and lower masses of the stars, which translate into larger transit
depths and Doppler signals. Moreover, the detection and follow-up studies of terrestrial planets in the
habitable zone are facilitated by the fact that the same stellar insolation as that received by the Earth
is achieved at shorter orbital periods around M dwarfs. The increased frequency and greater depth
of transits also renders the atmospheres of these worlds more accessible to transmission and emission
spectroscopy, and atmospheric studies are eagerly anticipated with both the upcoming James Webb
Space Telescope (e.g. Morley et al. 2017) and the Extremely Large Ground-based Telescopes (e.g.
Rodler & L´opez-Morales 2014). For these reasons, the only terrestrial exoplanets whose atmospheres
can be spectroscopically studied in the near-future will be those that orbit nearby mid-to-late M
dwarfs. Considering stars with masses less than 0.3M(cid:12) and within 15 parsecs, there are only four
such stars known to host transiting planets: GJ 1214 (Charbonneau et al. 2009), GJ 1132 (Berta-
Thompson et al. 2015), TRAPPIST-1 (Gillon et al. 2017), and LHS 1140 (Dittmann et al. 2017b).
There are also an additional 7 non-transiting systems: Proxima (Anglada-Escud´e et al. 2016), Ross
128 (Bonfils et al. 2018a), YZ Cet (Astudillo-Defru et al. 2017a), GJ 273 and GJ 3323 (Astudillo-
Defru et al. 2017b), Wolf 1061 (Astudillo-Defru et al. 2017b; Wright et al. 2016), and Kapteyn's star
(Anglada-Escud´e et al. 2014).
The Kepler dichotomy refers to the observed excess of stars with single transiting planets compared
to the expectations based on geometry and the population of stars with multiple transiting planets
(Lissauer et al. 2011). Ballard & Johnson (2016) studied this effect for M dwarfs, and found they
could account for the observed Kepler population if they posited that half of the planetary systems
have on average 7 planets with small mutual inclinations, while the other half of planetary systems
have only a single close-in planet, or multiple planets with a large range of mutual inclinations.
Numerous compact systems of small planets orbiting M dwarfs are known. In fact, Muirhead et al.
A second terrestrial planet orbiting LHS 1140
3
(2015) found that 21+7−5% of the Kepler M dwarfs host multiple planets with orbital periods less
than 10 days.
Indeed, of the 11 planetary systems of nearby mid-to-late M dwarfs described in
the preceding paragraph, the majority have (or are suspected to have) more than one planet. Our
intensive campaign with HARPS to follow-up GJ 1132 recently revealed the presence of at least
one additional planet (Bonfils et al. 2018b), although it does not appear to transit (Dittmann et al.
2017a).
Dittmann et al. (2017b) announced the discovery of a terrestrial planet on a 24.7-day orbit in the
habitable zone of LHS 1140, with an estimated zero-albedo equilibrium temperature of 230 ± 20
K. Given the context above, LHS 1140 is a particularly attractive target to search for co-planar
planets orbiting interior to the known planet. Therefore, since the discovery of LHS 1140 b, we
have intensively monitored this star with MEarth and HARPS in the hope that we would uncover
additional transiting planets. A secondary science goal of these observations is to refine the estimate
of the density of LHS 1140 b, and hence to address whether the ratio of the iron core to the rocky
mantle is consistent with that of the Earth and terrestrial planets generally, given the small spread
in abundances of Fe, Mg, and Si among nearby (albeit Sun-like) stars (Bedell et al. 2018). LHS 1140
is a slowly rotating and inactive star, and hence there is no reason to expect that stellar photospheric
effects will set a limit to the precision with which the planetary mass may be determined. In this
paper, we present a substantially improved estimate for the density of LHS 1140 b, and report the
discovery of another small terrestrial planet, LHS 1140 c, on a 3.8-day orbit.
This paper is structured as follows. Section 2 summarizes our knowledge of the host star LHS 1140.
Section 3 provides an overview of the different instruments and surveys that were used to conduct
photometric and spectroscopic monitoring of LHS 1140. Section 4 outlines the process that led to the
discovery of the new terrestrial companion LHS 1140 c. Subsequently, we performed joint modeling
of the two planets to estimate their orbital parameters as described in Section 5. We conclude by
discussing the scientific implications of our findings in Section 6.
2. LHS 1140
LHS 1140 is a M4.5-type main-sequence red dwarf (Dittmann et al. 2017b). The most up-to-date
parallax estimate for LHS 1140 is π = 0.066700 ± 0.000067(cid:48)(cid:48) (Gaia Collaboration et al. 2018) which
translates to a distance of 14.993 ± 0.015 pc. We note that this distance is substantially larger than
the previous best estimate of 12.47 pc by Dittmann et al. (2017b). Using the mass-luminosity relation
(MLR) for main-sequence M dwarfs by Benedict et al. (2016) and the 2MASS K-band apparent
brightness KS = 8.821± 0.024, we obtain an estimate for the stellar mass of M(cid:63) = 0.179± 0.014 M(cid:12).
We calculated the uncertainty on the mass by propagating the errors in the parallax and KS values
and adding a 0.014M(cid:12) term in quadrature to account for the scatter in the MLR. Therefore, we find
that the uncertainty on the stellar mass is completely dominated by the scatter in the MLR.
We obtained an updated estimate for the radius of LHS 1140 from our transit models in Section 5.1.
In particular, the a/R(cid:63) ratio (where a is the orbital semi-major axis and R(cid:63) is the radius of the star)
can be obtained directly from the transit duration, and the semi-major axis a can be calculated using
Kepler's Third Law knowing the orbital period P and stellar mass M(cid:63). This also leads to a direct
constraint on the stellar density from the transit parameters alone via ρ(cid:63) ≈ 3π
, assuming a
circular orbit (Seager & Mall´en-Ornelas 2003). In particular, we obtain R(cid:63) = 0.2111 ± 0.0059 R(cid:12)
from the transit models for LHS 1140 b, and R(cid:63) = 0.2165 ± 0.0057 R(cid:12) from the models for LHS
(cid:16) a
(cid:17)3
GP 2
R(cid:63)
Ment et al.
4
1140 c. We average the two and use R(cid:63) = 0.2139 ± 0.0041 as the final estimate. This result is
also consistent with the mass-radius relation from long-baseline optical interferometry of single stars
(Boyajian et al. 2012) which yields R(cid:63) = 0.209 ± 0.011 R(cid:12).
Due to the change in the distance estimate to LHS 1140, the stellar luminosity L(cid:63) also needs to
be reevaluated. We obtained a new value for the stellar luminosity by combining several bolometric
correction estimates. In particular, the Leggett et al. (2001) relation between BCJ and I − K yields
L(cid:63) = 0.00434 L(cid:12) whereas the Mann et al. (2015) relation between BCK and V −J gives L(cid:63) = 0.00431
L(cid:12). Alternatively, we can interpolate the BCV sequence of Pecaut & Mamajek (2013) based on the
V − K color of the star to get L(cid:63) = 0.00459 L(cid:12). We adopt as our final estimate the mean and
standard deviation of the three estimates, obtaining L(cid:63) = 0.00441 ± 0.00013 L(cid:12). We also adopt
the solar values of L(cid:12)= (3.8270 ± 0.0014) · 1026 W and Mbol,(cid:12) = 4.7554 used by Pecaut & Mamajek
(2013). As a result, we can estimate the effective surface temperature from the Stefan-Boltzmann
law as Teff,(cid:63) = 3216 ± 39 K.
Dittmann et al. (2017b) used the near-infrared spectral features of LHS 1140 to establish the
metallicity of the star as [Fe/H] = -0.24 ± 0.10. They infer the age of LHS 1140 to be larger than 5
Gyr based on the lack of active Hα emission as well as the slow rotation of the star. We adopt these
values.
The parameters of LHS 1140 can be found in Table 1.
3. PHOTOMETRIC AND RADIAL VELOCITY DATA
3.1. MEarth
MEarth-South is a telescope array consisting of eight 40-cm telescopes at the Cerro Tololo Inter-
national Observatory (CTIO) in Chile. The telescopes are operated on a (nearly) automated basis
and take data on every clear night. The observational strategy and the data reduction process are
described in greater detail in (Irwin et al. 2015) and Dittmann et al. (2017a). We note that the
MEarth data are corrected for the effect of differential color extinction by the Earth's atmosphere.
The primary driver of differential color extinction in our data is the variability in the amount of
precipitable water vapor in the atmosphere over the course of a night. Because the MEarth targets
are mid-to-late M dwarfs and typically the reddest object in any observing field, our target stars
are more sensitive to changes in water vapor than the field reference stars. We correct this effect
by measuring a "common mode" for all of our target M dwarfs. The common mode is defined as
the average differential (i.e. relative flux) light curve of all M dwarfs currently being observed by
MEarth-South in time bins of 0.02 days (28.8 minutes). This correlated behavior serves as a good
proxy for the local change in precipitable water vapor on these time scales.
MEarth data are reduced in real time during the night. During the course of normal operations,
MEarth-South is able to identify potential transits in-progress and instead of proceeding to the next
star in its target list, it can automatically decide to halt the data collection of other targets in order
to collect additional high-cadence data around the star showing signs of a potential transit event.
Normal operations are then resumed if the event is deemed to be spurious or the flux from the star
has returned to its normal level. This MEarth "trigger" mode enables us to potentially confirm a
planetary transit in real time and follow the transit all the way through egress. For a more detailed
description of the MEarth trigger mode, see Berta et al. (2012).
A second terrestrial planet orbiting LHS 1140
5
MEarth-South has been in operation since January of 2014, and we have been observing LHS
1140 since the beginning of this survey. Since 2014 we have taken 28,382 observations of LHS 1140
with MEarth-South. The observations contain the discovery data of LHS 1140 b (Dittmann et al.
2017b), including the high-cadence follow-up observations of LHS 1140 b's transits. The majority of
these data, however, are standard monitoring observations of LHS 1140. Since October of 2015, the
monitoring of LHS 1140 is carried out with two telescopes, with several exposures per visit.
3.2. Spitzer
We observed four transits of LHS 1140 b as part of the Spitzer DDT program 13174 (PI: Dittmann).
These data were taken with the Infrared Array Camera (IRAC) at 4.5 µm. The goal of this program
is to better determine the parameters of LHS 1140 b and to probe the system for signs of a transiting
exomoon or transit timing variations (TTVs). A manuscript analyzing these 4 transits is currently
under preparation. Here, we present one of these transit observations, which fortuitously captured a
transit of both LHS 1140 b and LHS 1140 c.
Observations spanned from 18 April 2018 03:21:22 to 18 April 2018 09:27:59 UTC. We placed
LHS 1140 in the portion of the detector that is well-characterized for the purpose of obtaining high
precision light curves. We obtained the data in subarray mode and utilized a small 16×16 pixel area
of the detector for our observation. During the first 78 minutes of the observations, the centroid of
LHS 1140 wandered nearly 0.3 pixels (0.37") before settling into the "sweet-spot" for the rest of the
observations. In order to correct for intra-pixel sensitivity variations, we calibrate our data with the
pixel-level decorrelation algorithm which depends on stable pointing to be effective. We exclude this
portion of the data and only use data for which the target has settled onto the same portion of the
detector. Subsequently, 50% of the image centroids fall within 0.069 pixels of the median centroid
location and 95% of the image centroids fall within 0.150 pixels of the median centroid location. We
then carry out the data reduction as described in Dittmann et al. (2017a). After obtaining the pixel-
level coefficients for our observations, we apply these coefficients to the unbinned and unnormalized
data. We sum the values of these weighted pixels in order to obtain the total flux from LHS 1140. We
apply a new outlier rejection criteria to these unbinned data. We reject all data points 7.5 median
absolute deviations from a 12-minute wide sliding median in order to eliminate single point outliers,
resulting in the rejection of 8 total data points (0.09% of the total number of data points).
Our final data set consists of 135 sets of 64 individual sub-array images, each with an integration
time of 2 seconds (our final datacube does not contain a full set of 64 images, as it was the last
datacube of the observing program), for a total of 8545 data points. These data were calibrated with
the Spitzer pipeline version S19.2.0 and the timestamps of each data point are calculated at the Solar
System barycenter in the TDB time system (Eastman et al. 2010).
3.3. HARPS
We intensively monitored LHS 1140 with the High Accuracy Radial velocity Planet Searcher
(HARPS) in the frame of the M dwarf program (ESO Id: 191.C-0873(A), 198.C-0838(A); PI: Bonfils)
and a survey dedicated to LHS 1140 (ESO Id: 0100.C-0884(A); PI: Astudillo-Defru). HARPS is a
fiber-fed echelle spectrograph located at the 3.6m telescope at the La Silla observatory in Chile. Its
resolving power is R=115,000, and it has a wavelength coverage between 380 nm and 690 nm over
72 orders (Mayor et al. 2003). HARPS achieves a sub-ms−1 long-term precision thanks to its tem-
perature and pressure controlled environment and a simultaneous wavelength calibration through
6
Ment et al.
a second fiber. We opted to place the calibration fiber on the sky to avoid contamination of the
science spectra from a calibration spectrum, notably in the region of the Ca II H&K lines. This
choice does not degrade our RV precision, because for a star with the brightness of LHS 1140, our
RV precision is dominated by the photon-noise of the stellar spectrum, as opposed to the calibration
of the spectrograph.
We obtained 294 spectra of LHS 1140 between November 23, 2015 and January 15, 2018, spanning
784 days. In general, HARPS observations consist of two spectra per night with an exposure time of
1,800 seconds each. On some nights, we observed LHS 1140 one or three times. We discarded one
observation (BJD: 2457686.6227) with an exposure time of only 1.8 sec, so our analysis was done
over 293 radial velocity measurements.
The HARPS Data Reduction Software (DRS, Lovis & Pepe 2007) automatically reduces the spectra
and computes the stellar radial velocity. The latter is done by cross-correlating spectra with a mask
designed to match the spectral lines. The DRS uses a mask containing the vast majority of the
absorption lines present in the spectrum of an M dwarf, however, a non-negligible amount of lines are
left out. To recover as much Doppler signal as possible, we follow the recipe described in Astudillo-
Defru et al. (2015, and references therein): the radial velocities derived by the DRS are used to
shift all spectra to a common reference frame, then the median is computed to obtain an enhanced
stellar spectrum (template). This template is Doppler shifted in a range of velocities to maximize its
likelihood with each spectrum, resulting in the set of radial velocities used hereafter.
4. DISCOVERY OF LHS 1140 c
4.1. Machine learning to identify transit candidates
Articles describing the use of artificial neural networks to classify ground-based data in the exoplanet
field have only recently appeared in the literature (known examples include Dittmann et al. (2017b)
and Armstrong et al. (2018)). Nevertheless, given that the first planet in the LHS 1140 system
was discovered by Dittmann et al. (2017b) via a machine learning process, we decided to use a
similar approach to search for additional potential transit events in the MEarth data. MEarth
generates many "triggers" per night (see Section 3.1), most of which can be explained by variations
in the atmosphere (clouds, changes in the column density of precipitable water vapour), rapid stellar
variability (e.g. stellar flares), or instrumental systematics. Such false signals can often be identified
by correlating the change in flux with a series of atmospheric and instrumental parameters such as the
common mode (defined in Section 3.1), the FWHM or pixel position of the target, or simultaneous
variations in the fluxes of background stars. A small minority of triggers correspond to genuine
transit events, which are not expected to be significantly correlated with any of these parameters.
Our machine learning approach seeks to take advantage of this distinction - that is, we would like to
eliminate as many triggers as possible based on systematics, and keep the rest as potential transit
candidates.
After testing several neural network designs including feedforward and recurrent neural networks
(see Murphy (2012) for more information about neural networks), we settled for a simple feedforward
neural network (FNN) consisting of an input layer, a single hidden layer, and an output node. The
network design employed in this paper is similar to the one used by Dittmann et al. (2017b) and
can be summarized as follows. The input layer was composed of 10 neurons for atmospheric and
instrumental parameters including the airmass, angle of the frame relative to the reference frame,
A second terrestrial planet orbiting LHS 1140
7
Figure 1. MEarth trigger event from August 14, 2016. The golden points represent individual measurements
whereas the green triangles are weighted averages in 10-minute time bins. The dashed line illustrates the
predicted transit of LHS 1140 c based on the results from joint photometric and RV modeling. According
to the neural network, this event had a low probability of generating a trigger, meaning that the change in
flux could not be explained by variations in the atmospheric and systematic parameters.
RMS of the fluxes of background stars, and the FWHM, ellipticity and pixel position of the target,
as well as the common mode with its time derivative and offset from a 3-hour median. The hidden
layer was made up of 10 neurons as well, each fully connected to every input node. Finally, the FNN
had a single output node representing the probability that the given input corresponds to a trigger.
The data set, which was randomly split up into a training set (90%) and a validation set (10%),
consisted of 1539 real triggers from MEarth between the years of 2014 and 2017. We complemented
this set with an equal number of non-triggers which were selected randomly from the regular, low-
cadence observations while ensuring that each star contributed the same amount of triggers and
non-triggers (however, the number of triggers contributed by each star varied considerably). After
training the network, we reclassified the 1539 real triggers, and selected the events where the FNN
generated the lowest probabilities of being an expected trigger (that is, caused by atmospheric or
instrumental effects) for further inspection. These events, unlike the majority of triggers, did not
exhibit a detectable correlation with the atmospheric and instrumental parameters that we used as
inputs.
In particular, we identified two triggers belonging to LHS 1140 with very low trigger probabilities:
one from 15 September 2014 UTC which was the basis of the LHS 1140 b discovery by Dittmann
et al. (2017b), and another trigger from 14 August 2016 UTC (illustrated in Figure 1) which did not
overlap with any of the predicted transits of planet b. The estimated probabilities for the two events
to be triggers were 23.2% and 24.5%, respectively, which placed them among the top 5% events with
the lowest trigger probabilities. The latter event was then selected as a potential transit candidate.
4.2. BLS search
8
Ment et al.
We also decided to carry out a phase-folded box model search in the photometry using the Box-
fitting Least Squares (BLS) algorithm (Kov´acs et al. 2002). In the case of LHS 1140, we tested over
1.6 million evenly spaced orbital frequencies corresponding to periods between 0.2 and 10 days. Our
frequency spacing ensured that the maximum shift in orbital phase between two consecutive tested
frequencies was always less than one half of the spacing between the time bins (we binned the observed
fluxes into 10-minute error-weighted intervals). After fitting a box model to each orbital frequency,
we imposed 4 selection criteria on the BLS spectrum to eliminate false detections. Firstly, we ignored
any periods close to half a day or any integer number of days to avoid detecting variability that can
be attributed to the Earth-based observing strategy. Secondly, we ignored any anti-transits (where
the fitted transit depth was negative) as well as transit depths less than 0.1%. We also required
that the significance of the fitted model at a given period be at least 90%, estimated by repeatedly
refitting the data after adding normally distributed random offsets to the measured fluxes according
to their respective uncertainties, and rejecting any outcomes that resulted in anti-transits. Finally,
we calculated ∆χ2, the difference in χ2 between the best-fit box model and a constant flux model,
for each orbital frequency. In order to eliminate signals that are dominated by observations from a
single night, we required that no night contribute more than 70% of the total ∆χ2-difference (Burke
et al. 2006). We then ordered the remaining orbital frequencies in decreasing order of ∆χ2.
After masking out the predicted transits of LHS 1140 b, we investigated the 16 highest ∆χ2-peaks
in the BLS spectrum. One of these orbital periods, P = 3.77797 days, produced a phase-folded
model that back-predicted a transit on August 14, 2016, with the ingress and egress times consistent
with the MEarth trigger described in Section 4.1. Overall, the in-transit observations spanned 35
different nights, with no single night contributing more than 19% to the total ∆χ2-difference. The
predicted transit depth from the box model was 0.192% which corresponds to an Earth-sized planet.
The phase-folded model is displayed in Figure 2.
4.3. A transit of LHS 1140 c in the Spitzer data
We show our calibrated Spitzer data in Figure 3. In this plot, we see the transit of LHS 1140b
followed by a second transit event of smaller depth and shorter duration. This second transit event
is not associated with shifts in the pixel position of the centroid of the target. The duration of this
event is approximately 74 minutes with a depth of 3 mmag.
The discovery of the Spitzer transit event (by J.A.D.) happened independently of identifying the
trigger or phase-folded transits of LHS 1140 c in the MEarth data (led by K.M.) and independently of
the HARPS radial velocity identification of LHS 1140 c (led by N.A-D.). We discuss the simultaneous
fitting of these transits with radial velocity observations in Section 5.1.
4.4. Periodogram analysis of the RVs
Using HARPS radial velocities, we unveiled the signal of LHS 1140 c independently of the photom-
etry. A Generalized Lomb-Scargle (GLS, Zechmeister & Kurster 2009) periodogram analysis of the
RVs reveals several interesting signals, shown in Figure 4. The most significant peak lies close to 25
days which corresponds to the orbital period of LHS 1140 b. After subtracting a Keplerian model
for LHS 1140 b from the RVs, a strong signal emerges near the stellar rotation period Prot = 131 ± 5
days as well as the 3.8-day period, matching the one found by the BLS analysis in Section 4.2. After
subtracting a two-planet model, we produced a periodogram displayed in the third panel of Figure 4,
exhibiting several low-frequency peaks at around 65 days and longer periods. At least two of these
A second terrestrial planet orbiting LHS 1140
9
Figure 2. Left: A phase-folded display of the MEarth photometric data folded at a period of P = 3.77797
days. The yellow points represent the trigger from August 14, 2016, and the grey shaded area represents the
proposed transits of LHS 1140 c. The red triangles are 10-minute binned averages. Right: A zoomed-in
version of the left-side plot illustrating the transits of LHS 1140 c. The dashed line illustrates a fitted box
model for the transits. The horizontal axis is rescaled to count the number of transit durations from the
beginning of the transit.
peaks, at 68 days and 132 days, can be attributed to the stellar rotation period Prot and its harmonic
Prot/2. Collections of starspots can often be successfully modeled as a series of no more than the
first three or four harmonics of Prot (Jeffers & Keller 2009; Boisse et al. 2011; Haywood 2015), such
as in the cases of Corot-7 (Queloz et al. 2009; Boisse et al. 2011) and HD 189733 (Boisse et al.
2011). Therefore, we proceeded to generate a model for stellar activity that consists of a sum of two
sinusoidal signals with periods initialized at Prot and Prot/2 (adding in further harmonics was not
necessary in this case). Subtracting this model from the residuals significantly reduces any residual
power in the low-frequency signals beyond 65 days (displayed in the bottom panel of Figure 4). Thus,
we see no evidence that the previously significant GLS peaks between 65 and 130 days were caused
by anything other than stellar activity.
5. PARAMETER ESTIMATION FOR LHS 1140 b AND c
5.1. A joint photometry and RV analysis
We fit a joint model of the Spitzer observations, MEarth-South observations, and the radial velocity
measurements from HARPS to simultaneously constrain the masses, radii, and orbital parameters of
both planets, as well as the radial velocity variability due to the stellar activity of LHS 1140. Our
Spitzer data are shown in Figure 3 and described in Section 3.2. We have observed one transit of
LHS 1140 c with Spitzer and we present one transit of LHS 1140 b with Spitzer here as well. The
remaining transits of LHS 1140 b from the Spitzer program will be the subject of a future publication.
We do not currently have a full transit observation of LHS 1140 c with MEarth-South, but we fit the
transit observations obtained on 35 individual transit nights (see section 4.2), including the night of
the trigger.
10
Ment et al.
Figure 3. Spitzer data of a transit of LHS 1140 b taken on April 18, 2018. Unbinned data are shown in
blue and a binned version of the light curve is shown in salmon. The transit of LHS 1140 b is visible as
the first dip in the plot. Following the egress of LHS 1140 b, we serendipitously observed a transit of LHS
1140 c, visible as the second smaller and shorter transit in the light curve. Members of our team recovered
this planet independently in the Spitzer and MEarth photometry as well as in the HARPS radial velocity
measurements.
We fit our photometric observations with the batman code (Kreidberg 2015). batman is an optimized
python implementation of the analytical model for transit light curves from Mandel & Agol (2002)
and its erratum. We model the radial velocity variations of LHS 1140 from the orbits of the two
planets (consistent with the transit ephemeris) and we adopt a simple sinusoidal model to describe
the radial velocity variations of LHS 1140 due to the activity on the star.
We initiate the physical parameters for LHS 1140 b with the values found in Dittmann et al. (2017b).
We adopt limb darkening coefficients from Claret et al. (2012) for a 3300 K star with log(g)= 5.0 in
a Cousins I filter. The Cousins I filter has a similar effective wavelength to the MEarth bandpass.
For the Spitzer transits, we adopt limb darkening coefficients from Claret et al. (2013), which are
calculated for the Spitzer bandpass. We initiate the period of LHS 1140 c at the best fit period in
the phase-folded detection and a T0 estimated from the midpoint of the Spitzer transit. The radial
velocity amplitude for LHS 1140 c is initialized to an amplitude of 3 m s−1, which is the approximate
amplitude a rocky composition planet would exhibit at this orbital period. In order to efficiently
explore parameter space, we use the emcee code (Foreman-Mackey et al. 2013). emcee is a python
implementation of the Affine-Invariant Markov Chain Monte Carlo sampler (Goodman & Weare
2010). Each model is initiated with 100 walkers in a Gaussian ball located at this initial solution.
0.7000.7250.7500.7750.8000.8250.8500.875Time (BJDTDB)+2.458226e60.970.980.991.001.011.021.03Relative FluxA second terrestrial planet orbiting LHS 1140
11
Figure 4. A Generalized Lomb-Scargle (GLS) periodogram of the radial velocities. The top panel shows the
GLS for the original RV data set, whereas the second and the third panels illustrate the GLS of the residuals
after removing one or both planets, respectively. The residuals of the two-planet model exhibit a series of
significant GLS peaks at periods of 65 days and higher. However, they can be removed by introducing a
model for stellar activity which consists of the first two harmonics of the stellar rotation period, as shown
in the two bottom panels. As a result, we associate these peaks with stellar activity. Vertical dashed lines
are overplotted on all panels to represent the periods of every modeled signal (both planetary and activity-
induced). Horizontal lines estimate the 1% and 5% false alarm probabilities (FAP) to eliminate any peaks
arising from window functions.
We run each chain for 100,000 steps and discard the first 10% of steps so that the solution may
"burn-in" independent of the choice of initialization. We show the results of this analysis in Table 2.
We show the results of these fits in Figures 5 and 6. We are able to recover the phased transits
of LHS 1140 c in the MEarth-South photometry and simultaneously fit the photometry and radial
velocities for both planets. Our planetary parameters for LHS 1140 b are consistent with, but more
precise than, those presented in Dittmann et al. (2017b). We find that LHS 1140 c has a transit depth
of 3.0 mmag. We present further discussion of the results of these fits and the physical parameters
of the LHS 1140 system in Section 6.
Up to this point, we have modeled the radial velocities using a simplistic model: we have only
considered the effects of both of the planets in the system and have modeled the stellar signal as
a single sinusoid at the stellar rotation period (measured through our photometry). This analysis
assumes that each data point is independent and that there is no correlated noise on any timescale.
12
Ment et al.
Parameter
Stellar parameters
Right ascension (J2000)
Declination (J2000)
Proper motion (mas yr−1)
Apparent brightness (mag)
Distance (pc)
Mass (M(cid:12))
Radius (R(cid:12))
Luminosity (L(cid:12))
Effective temperature (K)
Metallicity [Fe/H]
Age (Gyr)
Rotational period (days)
Table 1. System parameters for LHS 1140
Values for LHS 1140
Sourcea
00h 44min 59.3s
-15◦ 16' 18"
µα = 317.59 ± 0.12 µδ = −596.62 ± 0.09
VJ = 14.18 ± 0.03
J = 9.612 ± 0.023
RC = 12.88 ± 0.02 H = 9.092 ± 0.026
IC = 11.19 ± 0.02 KS = 8.821 ± 0.024
14.993 ± 0.015
0.179 ± 0.014
0.2139 ± 0.0041
0.00441 ± 0.00013
3216 ± 39
-0.24 ± 0.10
> 5
131 ± 5
(1)
(1)
(1)
(2)(4)
(1)
(3)
(3)
(3)
(3)
(4)
(4)
(4)
< 0.31
0.07390 ± 0.00008
LHS 1140 b
LHS 1140 c
4.85 ± 0.55
< 0.06
24.736959 ± 0.000080
3.777931 ± 0.000003
2.35 ± 0.49
2456915.71154 ± 0.00004
2458226.843169 ± 0.000026
89.89+0.05−0.03
95.34 ± 1.06
Parameter
Modeled transit and RV parameters
Orbital period P (days)
RV semi-amplitude K (m s−1)
Eccentricity e (90% confidence)
Time of mid-transit tT (BJD)
Inclination i (deg)
Planet-to-star radius ratio r/R(cid:63)
a/R(cid:63) ratio
Derived planetary parameters
Mass m (M⊕)
Radius r (R⊕)
Density ρ (g cm−3)
Surface gravity g (m s−2)
Semi-major axis a (AU)
Incident flux S (S⊕)
Equilibrium temperatureb Teq (K)
a(1) Gaia Collaboration et al. (2018), (2) Cutri et al. (2003), (3) This work, (4) Dittmann et al. (2017b).
b The equilibrium temperature assumes a Bond albedo of 0. For an albedo of AB, the reported temperature
needs to be multiplied by (1 − AB)1/4.
0.0936 ± 0.0024
0.503 ± 0.030
0.05486 ± 0.00013
89.92+0.06−0.09
26.57 ± 0.05
6.98 ± 0.89
1.727 ± 0.032
1.81 ± 0.39
1.282 ± 0.024
4.7 ± 1.1
10.6 ± 2.2
0.02675 ± 0.00070
6.16 ± 0.37
438 ± 9
7.5 ± 1.0
23.7 ± 2.7
235 ± 5
We now relax these assumptions to more accurately capture the correlated noise and errors present
in our radial velocity measurements.
5.2. Remodeling the RVs with Gaussian process regression
A second terrestrial planet orbiting LHS 1140
13
Table 2. Model parameters from the joint photometry and radial velocity fit in Section 5.1
Parameter
Explanation
Orbital period of LHS 1140 b
Transit time of LHS 1140 b
Ratio of LHS 1140 b and LHS 1140's radii
Ratio of semi-major axis to stellar radius for LHS 1140 b
Inclination angle of LHS 1140 b
Value
24.736959 ± 0.000080
2456915.71154 ± 0.000040
0.07390 ± 0.00008
95.34 ± 1.06
89.89+0.05−0.03
0 (fixed; < 0.094 at 99% confidence) Eccentricity of LHS 1140 ba
4.71 ± 0.09
3.777931+0.000004
−0.000002
2458226.843169 ± 0.000026
0.05486 ± 0.00013
26.57 ± 0.05
89.92+0.06−0.09
0 (fixed; < 0.236 at 99% confidence) Eccentricity of LHS 1140 ca
2.42 ± 0.10
Pb (days)
T0,b (BJD)
rb
r∗
ab
r∗
ib (degrees)
eb
Kb (m s−1)
Pc (days)
T0,c (BJD)
rc
r∗
ac
r∗
ic (degrees)
ec
Kc (m s−1)
γ (km s−1) −13.23851 ± 0.00008
K∗ (m s−1)
φ∗ (BJD)
P∗ (days)
a4.5
b4.5
aMEarth
bMEarth
aThese values were merely used as intermediate results, and were subsequently supplanted by the values from the
Gaussian Process modeling, listed in Table 1.
RV amplitude of LHS 1140 ba
Orbital period of LHS 1140 c
Transit time of LHS 1140 c
Ratio of LHS 1140 c and LHS 1140's radii
Ratio of semi-major axis to stellar radius for LHS 1140 c
Inclination angle of LHS 1140 c
RV amplitude of LHS 1140 ca
RV zeropoint of LHS 1140a
RV amplitude of stellar activitya
RV phase of stara
RV period of stara
Spitzer limb darkening coefficient
Spitzer limb darkening coefficient
MEarth limb darkening coefficient
MEarth limb darkening coefficient
3.0 ± 0.1
2457766.66 ± 0.68
132.26 ± 0.37
0.0104 ± 0.0007
0.2175+0.0012
−0.0042
0.195+0.022
−0.007
0.356+0.013
−0.11
Figure 5. Left: MEarth-South photometry of LHS 1140 phased to the orbit of LHS 1140 c. The salmon
line is our joint photometry and radial velocity fit. Here we have removed all transits of LHS 1140 b. Right:
Binned Spitzer photometry of the double transit of LHS 1140 b (left dip) and LHS 1140 c (right dip). The
salmon line is our joint model fit for the system.
0.4000.4250.4500.4750.5000.5250.5500.5750.600Phase of LHS 1140c (transit at 0.5)0.9800.9850.9900.9951.0001.0051.0101.0151.020Relative Flux0.700.750.800.850.90Time (BJDTDB)+2.458226e60.9900.9920.9940.9960.9981.0001.0021.004Relative Flux14
Ment et al.
Figure 6. Radial velocity measurements phased to the orbit of LHS 1140 b (left) and LHS 1140 c (right).
Data points are in blue with the model radial velocity curve shown in salmon. For each curve we have
removed the effects of stellar radial velocity and the radial velocity induced by the other planet. We find no
evidence for eccentricity in either planet.
As demonstrated by the GLS periodogram in Figure 4, the radial velocities exhibit significant cor-
relation near the stellar rotation period of 131± 5 days as well as its harmonics. The exact properties
of these quasi-periodic variations in M dwarfs are yet to be resolved, although they are commonly
believed to be linked to features on the stellar surface such as dark spots and granulation. Overall,
a constant sinusoidal model (as presumed in Section 5.1) might be an inaccurate representation of
activity-induced RV modulations. In pursuit of a more accurate accounting of the latter, we employed
Gaussian process (GP) regression (Rasmussen & Williams 2006) which has recently been applied to
several exoplanetary systems including Corot-7 (Haywood et al. 2014; Faria et al. 2016), Kepler-21
(L´opez-Morales et al. 2016), Kepler-1655 (Haywood et al. 2018), K-2 18 (Cloutier et al. 2017), Alpha
Centauri B (Rajpaul et al. 2015), and LHS 1140 (Dittmann et al. 2017b). Our goal in using GP
regression is to make as few assumptions about the exact form of the signal due to stellar activity
as possible. We did not include the photometric data in our GP framework in order to make the
analysis computationally manageable. Our GP regression was implemented by using a quasi-periodic
kernel of the form:
−(t − t(cid:48))2
η2
2
− sin2(cid:16) π(t−t(cid:48))
η3
(cid:17)
k(t, t(cid:48)) = η2
1 · exp
η2
4
(1)
where η1 represents the amplitude of the covariance function, η2 is the evolution timescale of activity-
inducing stellar surface features, η3 is the recurrence timescale (the expected period of correlation),
and η4 determines the amount of high-frequency structure in the GP. The Gaussian process framework
is extremely versatile, and therefore it is often desirable to constrain the values of the hyperparameters
with tight priors based on physical intuition about the system. We decided to adopt our priors from
Dittmann et al. (2017b) for consistency purposes. As a result, the recurrence timescale η3 was
constrained by a Gaussian prior centered at the stellar rotation period (131 ± 5 days). The active
regions of the star are likely to remain stable over a few rotation periods, so we adopt a Gaussian
prior at three times the rotation period (393 ± 30 days) for the evolution timescale η2. Similarly
to Dittmann et al. (2017b), choosing a different value of the same order of magnitude does not
significantly change our results. The structure parameter η4 is constrained by a Gaussian prior
0.00.20.40.60.81.0Phase (relative to b)15105051015Radial Velocity (m/s)0.00.20.40.60.81.0Phase (relative to c)151050510Radial Velocity (m/s)A second terrestrial planet orbiting LHS 1140
15
centered at 0.5 ± 0.05 which allows for two or three activity-induced peaks in the RV curve per
rotation (L´opez-Morales et al. 2016; Haywood et al. 2018). RV models with more than three peaks
per rotation would likely be unphysical due to the limb darkening and foreshortening effects that
effectively erase any structure in the photometric and RV signatures (Jeffers & Keller 2009).
We represented each planet i by a Keplerian model with an amplitude Ki, an orbital period Pi,
and a time of transit tT,i. The amplitudes Ki (as well as η1) were constrained only by a modified
Jeffreys prior. In order to accommodate the results of the joint fit from Section 5.1, we implemented
tight Gaussian priors on Pi and tT,i centered at the joint fit results while setting the widths equal
to the respective uncertainties. Initially, our models also included eccentricities ei and arguments of
periapsis ωi. However, we found the improvement in likelihood statistically negligible: the Bayesian
information criterion (BIC) increased from 1173 to 1218 when substituting a two-planet eccentric
model for a circular one. Thus, we used the eccentric model only to place an upper limit on the
orbital eccentricities: we optimized the rest of the parameters assuming e = 0. The parameters were
optimized via an affine-invariant Markov Chain Monte Carlo (MCMC) routine (Goodman & Weare
2010). We present a list of all modeled parameters with their appropriate prior distributions in Table
3.
Our best fit yielded an amplitude of Kb = 4.85±0.55 m s−1 and a period of Pb = 24.73712±0.00032
days for planet b. Similarly for planet c, we obtain Kc = 2.35 ± 0.49 m s−1and Pc = 3.777950 ±
0.000007 days. Using these results, we can estimate the planetary masses as mb = 6.98±0.89 M⊕ and
mc = 1.81±0.39 M⊕ (including the propagated uncertainties from the stellar mass). We also obtained
a GP amplitude of η1 = 2.68± 0.81 m s−1, GP timescales of η2 = 400± 38 days and η3 = 134.8± 3.2
days, and a structure parameter value of η4 = 0.51 ± 0.06. The latter three are all consistent with
our initial priors. Using the eccentric models, we can place upper limits of eb < 0.060 and ec < 0.313
on the orbital eccentricities with 90% confidence. However, we note that the posterior probability
distribution for the eccentricity of planet c is significantly lopsided, with the 80% confidence limit
at ec < 0.14. We decided to retain the orbital periods and times of transit from the combined fit in
Section 5.1 since these are constrained much more tightly by photometry. The uncertainties on all
of the parameters were obtained by calculating the 14th and 86th percentiles of the final posterior
distributions, and the best-fit values were taken from the medians of the PDFs.
Our best circular RV model is illustrated in Figure 7, and our final values for all modeled and
derived parameters are given in Table 1. We note that the 1σ uncertainties for Kb and Kc from
our updated RV solution are now consistent with the expected uncertainties based on the Fisher
information. Previously, Cloutier et al. (2018) found that the measurement precision of Kb in the
2 times larger than its theoretical value based
1-planet model of Dittmann et al. (2017b) was about
on the number of RV data points.
√
6. DISCUSSION AND CONCLUSION
Dittmann et al. (2017b) first announced the discovery of a transiting super-Earth orbiting the
nearby M dwarf LHS 1140. We managed to better constrain the properties of the planet by estimating
its mass mb = 6.98 ± 0.89 M⊕ and radius rb = 1.727 ± 0.032 R⊕. These improved values are
significantly influenced by the revised stellar mass and radius from more accurate measurements
as well as the new parallax value for LHS 1140 from Gaia DR2. Furthermore, we estimate the
eccentricity of LHS 1140 b to be below 0.06 with a 90% confidence, consistent with a circular or an
Earth-like orbit.
16
Ment et al.
Table 3. Modeled parameters and their prior distributions
Parameter
Prior distribution
Gaussian (24.736959, 0.000080), (3.777931, 0.000003)
Uniform [0, 1)
Uniform [0, 360)
Gaussian (2456915.71154, 0.00004), (2458226.843169, 0.000026)
P (days)
K (m s−1) Modified Jeffreys (σRV)
e
ω (deg)
tT (BJD)
η1 (m s−1) Modified Jeffreys (σRV)
η2 (days)
η3 (days)
η4
RV0
Gaussian (393, 30)
Gaussian (131, 5)
Gaussian (0.5, 0.05)
Uniform (−∞, +∞)
We also see strong evidence of a second terrestrial planet on a Pc = 3.777950 ± 0.000005 day
orbit around LHS 1140. In particular, we managed to observe several transits of LHS 1140 c with the
ground-based MEarth survey as well as a single transit in April 2018 using the Spitzer Space Telescope.
The planetary signal was also discovered independently in the radial velocities from the HARPS
M dwarfs survey. Combining photometric and RV data yields a mass estimate of mc = 1.81 ± 0.39
M⊕ and a planetary radius of rc = 1.282 ± 0.024 R⊕. The orbital eccentricity is ec < 0.14 with 80%
confidence and ec < 0.31 with 90% confidence.
Based on our mass and radius estimates for LHS 1140 b and c, we can calculate the mean densities
of the two planets: ρb = 7.5 ± 1.0 g cm−3 and ρc = 4.7 ± 1.1 g cm−3. We note that the density for
planet b decreased from the ρb = 12.5 ± 3.4 g cm−3 reported by Dittmann et al. (2017b), mostly due
to the 21% increase in the planetary radius. Both planets are consistent with a rocky composition, as
can be seen from the mass-radius diagram in Figure 8 which also illustrates the expected densities for
several theoretical bulk compositions (Zeng et al. 2016). The density of LHS 1140 c is also consistent
with the findings of Rogers (2015) and Dressing et al. (2015) who observed that planets with radii
below ∼1.6 M⊕ are preferentially rocky (that is, heavy enough to be composed of iron and silicates
alone) whereas larger planets tend to have lower densities which implies voluminous layers of volatiles
(H/He and astrophysical ices). LHS 1140 b, on the other hand, is consistent with a rocky composition
despite having a radius of 1.73 ± 0.03 R⊕. Therefore, we see further evidence that the transition
between planets with and without volatile envelopes is gradual (Rogers 2015). We note that LHS
1140 b falls at the minimum of the occurrence rate versus planet radius recently presented by Fulton
et al. (2017).
We can also estimate the equilibrium black-body temperatures of the two planets by using the
stellar temperature and Stefan-Boltzmann law. In particular, we obtain Teq,b = 235 ± 5 K and Teq,c
= 438 ± 9 K for planets b and c, respectively, assuming a Bond albedo of 0. An Earth-like albedo of
0.3 would yield temperatures of 215 K and 401 K, respectively. The incident fluxes on the two planets
compared to Earth are Sb = 0.503 ± 0.030 S⊕ and Sc = 6.16 ± 0.37 S⊕. This implies that LHS 1140
c is hot enough to be a Mercury analogue and is likely not habitable in the Earth-based sense. It
is also likely to be tidally locked and, barring gravitational perturbations from additional planetary
companions, is expected to be in spin-orbit resonance. The outer planet (LHS 1140 b), however,
A second terrestrial planet orbiting LHS 1140
17
Figure 7. The best 2-planet model (with circular orbits) for the radial velocities. The top panels show the
full model (and the corresponding RV residuals) including the two planets as well as an activity-induced
offset from Gaussian Process (GP) regression. The bottom panels show the signals for each planet separately,
with the other planet as well as the GP-predicted offset subtracted out, folded at the appropriate orbital
periods.
receives half the flux incident on Earth, and could thereby be more accurately characterized as a
Super-Mars. Depending on the atmospheric properties, the conservative inner and outer habitable
zone limits for LHS 1140 would be 0.07 AU and 0.14 AU, respectively (Kopparapu et al. 2013),
placing LHS 1140 b well within the conservative habitable zone. This is mostly consistent with the
findings of Kane (2018), although we place LHS 1140 b deeper within the habitable zone due to our
revised values for the orbital semi-major axis as well as the effective stellar temperature.
We note that the orbital period ratio of 1:6.55 between the two planets means that they are unlikely
to be in direct orbital resonance. Orbital resonances often indicate a history of planetary migration,
such as in the four-planet Kepler-223 system (Mills et al. 2016) or around TRAPPIST-1 (Tamayo
et al. 2017). However, recent simulations suggest that most hot super-Earths like LHS 1140 c are
18
Ment et al.
Figure 8. Mass-radius distribution of known nearby small exoplanets orbiting M dwarfs (distances less than
15 pc and host masses less than 0.3 M(cid:12)). The solid and the dashed lines show the expected mean densities
for various planetary compositions (Zeng et al. 2016). The masses and radii for the TRAPPIST-1 planets
were adopted from Grimm et al. (2018). LHS 1140 b and c are both consistent with a rocky (predominantly
iron and magnesium silicates) composition.
not expected to form in situ (Cossou et al. 2014; Ogihara et al. 2015), although resonant chains
can also be disrupted by the dispersal of the gaseous disk (Cossou et al. 2014). Nonetheless, the
relatively large difference between the orbital periods increases the likelihood of additional unseen
companions, especially since many known planetary systems with multiple close-in super-Earths are
much more compact. This view is also supported by the Kepler dichotomy (Lissauer et al. 2011;
Ballard & Johnson 2016). In particular, M dwarfs seem to exhibit either a single transiting planet
(explained by a single planet or a multiple-planet system with large mutual orbital inclinations) or
a large number of near-coplanar transiting planets (Ballard & Johnson 2016). The similar orbital
inclinations of LHS 1140 b and c (89.89 and 89.92 degrees, respectively) suggest that LHS 1140 might
belong to the latter group, with more planets yet to be discovered. Despite the fact that we do not
currently see additional significant signals in the photometry and the radial velocities (as illustrated
by the GLS periodogram in Figure 4), we strongly encourage further photometric monitoring as well
as high-precision RV follow-up. We also note that LHS 1140 is one of the few nearby systems where
the masses and radii of the planets can be determined with a relatively high accuracy, due to our
comparatively good knowledge of the stellar parameters.
A second terrestrial planet orbiting LHS 1140
19
Planets in the LHS 1140 system are also potential targets for atmospheric spectroscopy observations
with the James Webb Space Telescope (JWST). Morley et al. (2017) modeled the emission and
transmission spectra of LHS 1140 b for varying surface pressures and atmospheric compositions,
and noted that a 5σ detection of transmission spectral features would likely be very challenging,
if the atmosphere had a similar mean molecular weight compared to that of the Earth. However,
both emission and transmission spectra are sensitive to the molecular composition of the atmosphere,
which can strongly depend on temperature; the emission spectra are directly sensitive to temperature
as well. Due to its much higher equilibrium temperature and lower surface gravity, LHS 1140 c
would be a significantly better candidate for the detection of both emission and transmission spectral
features. In particular, the two planets tested by Morley et al. (2017) that have radii and temperatures
similar to those of LHS 1140 c - namely TRAPPIST-1 b and GJ 1132 b - would likely require less
than a dozen transits or eclipses to detect a Venus-like atmosphere, depending on the Bond albedo
and surface pressure. Comparative planetology studies are often challenging due to the fact that
planets in different systems evolve in very different stellar environments. Therefore, characterizing
the atmospheres of multiple planets in the same planetary system would likely yield a much better
understanding of how planets and their surface environments form and develop.
The MEarth team acknowledges funding from the David and Lucile Packard Fellowship for Science
and Engineering (awarded to D.C.). This material is based on work supported by the National Science
Foundation under grants AST-0807690, AST-1109468, AST-1004488 (Alan T. Waterman Award) and
AST-1616624. This publication was made possible through the support of a grant from the John
Templeton Foundation. The opinions expressed in this publication are those of the authors and do
not necessarily reflect the views of the John Templeton Foundation. This work was performed in part
under contract with the Jet Propulsion Laboratory (JPL) funded by NASA through the Sagan Fel-
lowship Program executed by the NASA Exoplanet Science Institute (in support of R.D.H.). J.A.D.
acknowledges support by the Heising-Simons Foundation as a 51 Pegasi b postdoctoral fellow. N.A-
D. acknowledges support from FONDECYT 3180063. X.B., J.M-A., and A.W. acknowledge funding
from the European Research Council under the ERC Grant Agreement n. 337591-ExTrA. R.C. ac-
knowledges support from the Natural Sciences and Engineering Research Council of Canada. E.R.N.
is supported by an NSF Astronomy and Astrophysics Postdoctoral Fellowship under award AST-
1602597. N.S.C. acknowledges support from FEDER - Fundo Europeu de Desenvolvimento Regional
funds through the COMPETE 2020 - Programa Operacional Competitividade e Internacionaliza¸cao
(POCI), and Portuguese funds through FCT - Funda¸cao para a Ciencia e a Tecnologia in the frame-
work of the project POCI-01-0145-FEDER-028953 and POCI-01-0145-FEDER-032113. We further
acknowledge the support from FCT through national funds and by FEDER through COMPETE2020
in the form of grant UID/FIS/04434/2013 & POCI-01-0145-FEDER-007672. This publication makes
use of data products from the Two Micron All Sky Survey (2MASS), which is a joint project of the
University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of
Technology, funded by NASA and the National Science Foundation. This publication makes use of
data products from the Wide-field Infrared Survey Explorer, which is a joint project of the University
of California, Los Angeles, and the JPL/California Institute of Technology, funded by NASA. This
research has made extensive use of the NASA Astrophysics Data System (ADS), and the SIMBAD
database, operated at CDS, Strasbourg, France.
20
Ment et al.
Software: batman (Kreidberg 2015), HARPS Data Reduction Software (Lovis & Pepe 2007), emcee
(Foreman-Mackey et al. 2013)
REFERENCES
Anglada-Escud´e, G., Arriagada, P., Tuomi, M.,
Boyajian, T. S., von Braun, K., van Belle, G.,
et al. 2014, MNRAS, 443, L89,
doi: 10.1093/mnrasl/slu076
Anglada-Escud´e, G., Amado, P. J., Barnes, J.,
et al. 2016, Nature, 536, 437,
doi: 10.1038/nature19106
Armstrong, D. J., Gunther, M. N., McCormac, J.,
et al. 2018, MNRAS,
doi: 10.1093/mnras/sty1313
Astudillo-Defru, N., Bonfils, X., Delfosse, X.,
et al. 2015, A&A, 575, A119,
doi: 10.1051/0004-6361/201424253
Astudillo-Defru, N., D´ıaz, R. F., Bonfils, X., et al.
2017a, A&A, 605, L11,
doi: 10.1051/0004-6361/201731581
Astudillo-Defru, N., Forveille, T., Bonfils, X.,
et al. 2017b, A&A, 602, A88,
doi: 10.1051/0004-6361/201630153
Ballard, S., & Johnson, J. A. 2016, ApJ, 816, 66,
doi: 10.3847/0004-637X/816/2/66
et al. 2012, ApJ, 757, 112,
doi: 10.1088/0004-637X/757/2/112
Burke, C. J., Gaudi, B. S., DePoy, D. L., & Pogge,
R. W. 2006, AJ, 132, 210, doi: 10.1086/504468
Charbonneau, D., Berta, Z. K., Irwin, J., et al.
2009, Nature, 462, 891,
doi: 10.1038/nature08679
Claret, A., Hauschildt, P. H., & Witte, S. 2012,
A&A, 546, A14,
doi: 10.1051/0004-6361/201219849
-- . 2013, A&A, 552, A16,
doi: 10.1051/0004-6361/201220942
Cloutier, R., Doyon, R., Bouchy, F., & H´ebrard,
G. 2018, ArXiv e-prints, arXiv:1807.01263.
https://arxiv.org/abs/1807.01263
Cloutier, R., Astudillo-Defru, N., Doyon, R., et al.
2017, A&A, 608, A35,
doi: 10.1051/0004-6361/201731558
Bedell, M., Bean, J. L., Melendez, J., et al. 2018,
Cossou, C., Raymond, S. N., Hersant, F., &
ArXiv e-prints.
https://arxiv.org/abs/1802.02576
Pierens, A. 2014, A&A, 569, A56,
doi: 10.1051/0004-6361/201424157
Benedict, G. F., Henry, T. J., Franz, O. G., et al.
Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al.
2016, AJ, 152, 141,
doi: 10.3847/0004-6256/152/5/141
Berta, Z. K., Irwin, J., Charbonneau, D., Burke,
C. J., & Falco, E. E. 2012, AJ, 144, 145,
doi: 10.1088/0004-6256/144/5/145
Berta-Thompson, Z. K., Irwin, J., Charbonneau,
D., et al. 2015, Nature, 527, 204,
doi: 10.1038/nature15762
Boisse, I., Bouchy, F., H´ebrard, G., et al. 2011,
A&A, 528, A4,
doi: 10.1051/0004-6361/201014354
Bonfils, X., Delfosse, X., Udry, S., et al. 2013,
A&A, 549, A109,
doi: 10.1051/0004-6361/201014704
Bonfils, X., Astudillo-Defru, N., D´ıaz, R., et al.
2018a, A&A, 613, A25,
doi: 10.1051/0004-6361/201731973
2003, VizieR Online Data Catalog, II/246
Dittmann, J. A., Irwin, J. M., Charbonneau, D.,
Berta-Thompson, Z. K., & Newton, E. R. 2017a,
AJ, 154, 142, doi: 10.3847/1538-3881/aa855b
Dittmann, J. A., Irwin, J. M., Charbonneau, D.,
et al. 2017b, Nature, 544, 333,
doi: 10.1038/nature22055
Dressing, C. D., & Charbonneau, D. 2013, ApJ,
767, 95, doi: 10.1088/0004-637X/767/1/95
-- . 2015, ApJ, 807, 45,
doi: 10.1088/0004-637X/807/1/45
Dressing, C. D., Charbonneau, D., Dumusque, X.,
et al. 2015, ApJ, 800, 135,
doi: 10.1088/0004-637X/800/2/135
Eastman, J., Siverd, R., & Gaudi, B. S. 2010,
PASP, 122, 935, doi: 10.1086/655938
Bonfils, X., Almenara, J. M., Cloutier, R., et al.
Faria, J. P., Haywood, R. D., Brewer, B. J., et al.
2018b, ArXiv e-prints, arXiv:1806.03870.
https://arxiv.org/abs/1806.03870
2016, A&A, 588, A31,
doi: 10.1051/0004-6361/201527899
A second terrestrial planet orbiting LHS 1140
21
Foreman-Mackey, D., Hogg, D. W., Lang, D., &
Lissauer, J. J., Ragozzine, D., Fabrycky, D. C.,
Goodman, J. 2013, PASP, 125, 306,
doi: 10.1086/670067
Fulton, B. J., Petigura, E. A., Howard, A. W.,
et al. 2017, AJ, 154, 109,
doi: 10.3847/1538-3881/aa80eb
Gaia Collaboration, Brown, A. G. A., Vallenari,
A., et al. 2018, ArXiv e-prints.
https://arxiv.org/abs/1804.09365
Gillon, M., Triaud, A. H. M. J., Demory, B.-O.,
et al. 2017, Nature, 542, 456,
doi: 10.1038/nature21360
Goodman, J., & Weare, J. 2010, Communications
in Applied Mathematics and Computational
Science, Vol. 5, No. 1, p. 65-80, 2010, 5, 65,
doi: 10.2140/camcos.2010.5.65
Grimm, S. L., Demory, B.-O., Gillon, M., et al.
2018, A&A, 613, A68,
doi: 10.1051/0004-6361/201732233
Haywood, R. D. 2015, PhD thesis, University of
St Andrews
Haywood, R. D., Collier Cameron, A., Queloz, D.,
et al. 2014, MNRAS, 443, 2517,
doi: 10.1093/mnras/stu1320
et al. 2011, The Astrophysical Journal
Supplement Series, 197, 8,
doi: 10.1088/0067-0049/197/1/8
L´opez-Morales, M., Haywood, R. D., Coughlin,
J. L., et al. 2016, AJ, 152, 204,
doi: 10.3847/0004-6256/152/6/204
Lovis, C., & Pepe, F. 2007, A&A, 468, 1115,
doi: 10.1051/0004-6361:20077249
Mandel, K., & Agol, E. 2002, ApJL, 580, L171,
doi: 10.1086/345520
Mann, A. W., Feiden, G. A., Gaidos, E., Boyajian,
T., & von Braun, K. 2015, ApJ, 804, 64,
doi: 10.1088/0004-637X/804/1/64
Mayor, M., Pepe, F., Queloz, D., et al. 2003, The
Messenger, 114, 20
Mills, S. M., Fabrycky, D. C., Migaszewski, C.,
et al. 2016, Nature, 533, 509,
doi: 10.1038/nature17445
Morley, C. V., Kreidberg, L., Rustamkulov, Z.,
Robinson, T., & Fortney, J. J. 2017, ApJ, 850,
121, doi: 10.3847/1538-4357/aa927b
Muirhead, P. S., Mann, A. W., Vanderburg, A.,
et al. 2015, ApJ, 801, 18,
doi: 10.1088/0004-637X/801/1/18
Haywood, R. D., Vanderburg, A., Mortier, A.,
Mulders, G. D., Pascucci, I., & Apai, D. 2015,
et al. 2018, AJ, 155, 203,
doi: 10.3847/1538-3881/aab8f3
Irwin, J. M., Berta-Thompson, Z. K.,
Charbonneau, D., et al. 2015, in Cambridge
Workshop on Cool Stars, Stellar Systems, and
the Sun, Vol. 18, Cambridge Workshop on Cool
Stars, Stellar Systems, and the Sun, ed. G. T.
van Belle & H. C. Harris, 767 -- 772
Jeffers, S. V., & Keller, C. U. 2009, in 15th
Cambridge Workshop on Cool Stars, Stellar
Systems, and the Sun, Vol. 1094, 664 -- 667
Kane, S. R. 2018, ApJ, 861, L21,
doi: 10.3847/2041-8213/aad094
Kopparapu, R. K., Ramirez, R., Kasting, J. F.,
et al. 2013, ApJ, 765, 131,
doi: 10.1088/0004-637X/765/2/131
Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A,
391, 369, doi: 10.1051/0004-6361:20020802
Kreidberg, L. 2015, PASP, 127, 1161,
doi: 10.1086/683602
ApJ, 814, 130,
doi: 10.1088/0004-637X/814/2/130
Murphy, K. P. 2012, Machine learning : a
probabilistic perspective, Adaptive computation
and machine learning (Cambridge, Mass.: MIT
Press)
Ogihara, M., Morbidelli, A., & Guillot, T. 2015,
A&A, 578, A36,
doi: 10.1051/0004-6361/201525884
Pecaut, M. J., & Mamajek, E. E. 2013, The
Astrophysical Journal Supplement Series, 208,
9, doi: 10.1088/0067-0049/208/1/9
Queloz, D., Bouchy, F., Moutou, C., et al. 2009,
A&A, 506, 303,
doi: 10.1051/0004-6361/200913096
Rajpaul, V., Aigrain, S., Osborne, M. A., Reece,
S., & Roberts, S. 2015, MNRAS, 452, 2269,
doi: 10.1093/mnras/stv1428
Rasmussen, C. E., & Williams, C. K. I. 2006,
Gaussian Processes for Machine Learning
Rodler, F., & L´opez-Morales, M. 2014, ApJ, 781,
Leggett, S. K., Allard, F., Geballe, T. R.,
Hauschildt, P. H., & Schweitzer, A. 2001, ApJ,
548, 908, doi: 10.1086/319020
54, doi: 10.1088/0004-637X/781/1/54
Rogers, L. A. 2015, ApJ, 801, 41,
doi: 10.1088/0004-637X/801/1/41
22
Ment et al.
Seager, S., & Mall´en-Ornelas, G. 2003, ApJ, 585,
Zechmeister, M., & Kurster, M. 2009, A&A, 496,
1038, doi: 10.1086/346105
Tamayo, D., Rein, H., Petrovich, C., & Murray,
N. 2017, ApJL, 840, L19,
doi: 10.3847/2041-8213/aa70ea
Wright, D. J., Wittenmyer, R. A., Tinney, C. G.,
Bentley, J. S., & Zhao, J. 2016, ApJL, 817, L20,
doi: 10.3847/2041-8205/817/2/L20
577, doi: 10.1051/0004-6361:200811296
Zeng, L., Sasselov, D. D., & Jacobsen, S. B. 2016,
ApJ, 819, 127,
doi: 10.3847/0004-637X/819/2/127
|
1106.2047 | 1 | 1106 | 2011-06-10T13:25:29 | Planetary Core Formation with Collisional Fragmentation and Atmosphere to Form Gas Giant Planets | [
"astro-ph.EP"
] | Massive planetary cores ($\sim 10$ Earth masses) trigger rapid gas accretion to form gas giant planets \rev{such as} Jupiter and Saturn. We investigate the core growth and the possibilities for cores to reach such a critical core mass. At the late stage, planetary cores grow through collisions with small planetesimals. Collisional fragmentation of planetesimals, which is induced by gravitational interaction with planetary cores, reduces the amount of planetesimals surrounding them, and thus the final core masses. Starting from small planetesimals that the fragmentation rapidly removes, less massive cores are formed. However, planetary cores acquire atmospheres that enlarge their collisional cross section before rapid gas accretion. Once planetary cores exceed about Mars mass, atmospheres significantly accelerate the growth of cores. We show that, taking into account the effects of fragmentation and atmosphere, initially large planetesimals enable formation of sufficiently massive cores. On the other hand, because the growth of cores is slow for large planetesimals, a massive disk is necessary for cores to grow enough within a disk lifetime. If the disk with 100\,km-sized initial planetesimals is 10 times as massive as the minimum mass solar nebula, planetary cores can exceed 10 Earth masses in the Jovian planet region ($>5\,$AU). | astro-ph.EP | astro-ph |
Planetary Core Formation with Collisional Fragmentation and
Atmosphere to Form Gas Giant Planets
Hiroshi Kobayashi1, Hidekazu Tanaka2, Alexander V. Krivov1
1 Astrophysical Institute and University Observatory, Friedrich Schiller University, Schillergaesschen 2-3,
2 Institute of Low Temperature Science, Hokkaido University, Kita-Ku Kita 19 Nishi 8, Sapporo 060-0819,
07745, Jena, GERMANY
JAPAN
[email protected]
ABSTRACT
Massive planetary cores (∼ 10 Earth masses) trigger rapid gas accretion to form gas giant
planets such as Jupiter and Saturn. We investigate the core growth and the possibilities for cores
to reach such a critical core mass. At the late stage, planetary cores grow through collisions with
small planetesimals. Collisional fragmentation of planetesimals, which is induced by gravitational
interaction with planetary cores, reduces the amount of planetesimals surrounding them, and
thus the final core masses. Starting from small planetesimals that the fragmentation rapidly
removes, less massive cores are formed. However, planetary cores acquire atmospheres that
enlarge their collisional cross section before rapid gas accretion. Once planetary cores exceed
about Mars mass, atmospheres significantly accelerate the growth of cores. We show that, taking
into account the effects of fragmentation and atmosphere, initially large planetesimals enable
formation of sufficiently massive cores. On the other hand, because the growth of cores is slow
for large planetesimals, a massive disk is necessary for cores to grow enough within a disk lifetime.
If the disk with 100 km-sized initial planetesimals is 10 times as massive as the minimum mass
solar nebula, planetary cores can exceed 10 Earth masses in the Jovian planet region (> 5 AU).
Subject headings: planet and satellites:formation
1.
Introduction
Gas giant planets such as Jupiter and Sat-
urn form in gaseous disks. In the core-accretion
model, the accretion of planetesimals produces
cores of giant planets. Once a core reaches a
critical mass ∼ 10 Earth masses, it can rapidly
accrete gas to form a gas giant planet (Mizuno
1980; Bodenheimer & Pollack 1986; Ikoma et al.
2000). Gas giants must form within the lifetime
of gaseous disks (. 10 Myr).
For km-sized or larger planetesimals, gravita-
tional focusing enhances their collisional cross sec-
tions, resulting in a high collision probability for
low relative velocities. Relative velocities of large
bodies are kept lower than those of small ones due
to dynamical friction. A combination of gravita-
tional focusing and dynamical friction brings rapid
growth of large bodies, which is referred to as run-
away growth (Wetherill & Stewart 1989). Eventu-
ally, the runaway growth generates a small popu-
lation of large bodies called planetary embryos.
Planetary embryos keep their orbital separations
and hence grow through collisions with surround-
ing remnant planetesimals more slowly than in the
runaway mode (Kokubo & Ida 1998). This regime
is called oligarchic growth.
Kobayashi et al. (2010) pointed out that the
oligarchic growth halts due to fragmentation of
planetesimals.
In the oligarchic growth, the rel-
ative velocities of planetesimals are controlled by
the viscous stirring of embryos and gas drag. As
embryos grow, the velocities of remnant planetes-
imals are increased so greatly that collisions be-
1
tween planetesimals become destructive.
Such
collisions eject numerous fragments, which col-
lide with each other to produce further smaller
bodies. Planetesimals are therefore ground down
through such successive collisions (collision cas-
cade). The random velocities of small bodies are
strongly damped by gas drag and thereby the colli-
sional cascade no longer occurs for fragments with
radii . 1 -- 10 m. In the end such fragments drift
inward due to gas drag and are lost around em-
bryos. The collisional cascade combined with the
loss of fragments reduces the solid surface density
and hence final embryo masses. Kobayashi et al.
(2010) showed that the final embryo masses are as
small as Mars mass for 1-100 km-sized planetesi-
mals in the minimum mass solar nebula (hereafter
MMSN; Hayashi 1981).
Large planetesimals,
which are relatively hard to be broken collision-
ally, and a massive disk produce massive final em-
bryos. However, collisional fragmentation makes
it difficult to form giant planets along the lines
of the core-accretion model; starting from 100 km-
sized planetesimals, planetary embryos can reach
the critical core mass for gas accretion only inside
3 -- 4 AU in a disk that is 10 times more massive
than the MMSN model.
The motion of fragments . 1 m is coupled with
gas. The drift timescale of such fragments are rel-
atively long. Kenyon & Bromley (2009) proposed
that embryos may accrete a large amount of such
fragments. However, the strong gas drag in the
Stokes regime is dominant for fragments . 100 m
and damps the relative velocities to halt collision
cascade at 1 -- 10 m as mentioned above. Therefore,
only a small amount of coupled bodies are pro-
duced and hence they hardly contribute to embryo
growth (Kobayashi et al. 2010).
Many authors have investigated embryo growth
with N -body,
statistical, and hybrid simula-
tions (Kokubo & Ida 1996, 1998, 2000, 2002;
Inaba et al. 1999, 2001, 2003; Weidenschilling et al.
1997; Weidenschilling 2005, 2008; Kenyon & Bromley
2004, 2008; Chambers 2006, 2008; Kobayashi et al.
2010). Although providing most accurate dynam-
ical results, N -body simulations have difficulty
in producing numerous fragments and following
their fate. The fragmentation effect on embryo
growth has thus not been treated in detail in spite
of its importance. Recently, Levison et al. (2010)
included fragment production in their N -body
2
simulation. However, it is still difficult to treat
fragment -- fragment collisions. Such successive col-
lisions are essential in the collision cascade (e.g.,
Kobayashi & Tanaka 2010). Therefore, statisti-
cal simulations are a better method to accurately
investigate planet formation with fragmentation.
In the statistical simulation, the collisional
mass evolution of bodies is calculated within a
"particle-in-a-box" approximation. Bodies have
horizontal and vertical components of random ve-
locity relative to a circular orbit that are deter-
mined by their eccentricities and inclinations, re-
spectively. These velocities are changed by gravi-
tational interactions between the bodies and hence
affected by their mass spectrum, while the colli-
sion rates between the bodies depend on the ve-
locities. Therefore, the coupled mass and velocity
evolution needs to be solved (Wetherill & Stewart
1993). While the statistical method has ad-
vantages,
is the inability to
track the individual positions of planetesimals.
However, progress in planetary dynamic theory
(Greenzweig & Lissauer 1992;
Ida & Nakazawa
1989; Ohtsuki 1999; Stewart & Ida 2000; Ohtsuki et al.
2002) has helped to overcome this problem.
For example, Greenzweig & Lissauer (1992) and
Ida & Nakazawa (1989) provided detailed ex-
pressions for the probability of collisions be-
tween planetesimals orbiting a central star, while
Stewart & Ida (2000) and Ohtsuki et al. (2002)
derived improved equations for calculating the
evolution of random planetesimal velocities caused
by gravitational
it has
been shown that the recently developed statis-
tical codes can describe some aspects of the plan-
etary accumulation processes with the same ac-
curacy as N -body simulations (Inaba et al. 2001;
Kobayashi et al. 2010).
its weak point
interactions.
Finally,
Since the timescale of collision cascade strongly
Collisional
affects the final mass of planetary embryos (Kobayashi et al.
fragmentation outcome models are es-
2010),
sential
for embryo growth.
frag-
mentation includes several uncertain parame-
ters. Kobayashi & Tanaka (2010) constructed
a simple fragmentation model which is consis-
tent with laboratory experiments (Fujiwara et al.
1977; Takagi et al. 1984; Holsapple 1993) and
hydrodynamical
(Benz & Asphaug
1999) and analytically clarified which parame-
ters are essential. They found that the mass de-
simulations
pletion due to collision cascades is sensitive to
the total ejecta mass yielded by a single colli-
sion, while it is almost independent of the mass
of the largest ejecta fragment and the size dis-
tribution of ejecta over a realistic parameter re-
gion. Furthermore, fragmenting collisions are sub-
divided into two types, catastrophic disruption
and cratering (erosive collision). Although some
studies neglected or underestimated the effect of
cratering (Dohnanyi 1969; Williams & Wetherill
1994; Wetherill & Stewart 1993; Inaba et al. 2003;
Bottke et al. 2005), Kobayashi & Tanaka showed
that cratering collisions make a dominant contri-
bution to the collision cascade.
A planetary embryo larger than ∼ 10−2 Earth
masses acquires a tenuous atmosphere of gas from
the disk. Fragments are captured by the atmo-
sphere even if they do not collide directly with
the embryo,
implying that the collisional cross
section of the embryo is enhanced. This ef-
fect advances the growth of Mars-mass or larger
embryos (Inaba & Ikoma 2003). Embryos with
the atmospheres accrete fragments prior to their
drift inward and can acquire more than 10 Earth
masses starting from 10km-sized planetesimals
(Inaba et al. 2003). However, erosive collisions
and initial planetesimal sizes strongly affect final
embryo masses (Kobayashi et al. 2010).
This paper investigates the embryo growth tak-
ing into account erosive collisions and embryo's
atmosphere. Although growing embryos may fall
into a central star due to the type I migration (e.g.,
Tanaka et al. 2002), we neglect the migration here.
We perform both analytical studies and statistic
simulations, which extend those of Kobayashi et
al. (2010) by including atmospheric enhancement
of embryo growth. The goal is to find out what de-
termines embryo growth and whether an embryo
can reach the critical core mass. We introduce
the theoretical model in Section 2 and derive final
embryo masses taking into account atmosphere in
Section 3. In Section 4, we check solutions for fi-
nal masses against the statistical simulations. Sec-
tions 5 and 6 contain a discussion and a summary
of our findings.
2. THEORETICAL MODEL
2.1. Disk Model
We introduce a power-law disk model for the
initial surface mass density of solids Σs,0 and gas
Σg,0 such that
Σs,0 = ficeΣ1(cid:16) a
Σgas,0 = fgasΣ1(cid:16) a
1AU(cid:17)−q
1AU(cid:17)−q
g cm−2,
g cm−2,
(1)
(2)
where a is a distance from a central star, Σ1 is
the reference surface density at 1 AU, and q is the
power-law index of the radial distribution. The
gas-dust ratio fgas = 240 (Hayashi 1981). The fac-
tor fice that represents the increase of solid density
by ice condensation beyond the snow line aice is
given by fice = 1 (a < aice) and 4.2 (a ≥ aice). In
the MMSN model, Σ1 = 7.1 g cm−2 and q = 3/2.
If the disk is optically thin,
aice = 2.7(cid:18) L∗
L⊙(cid:19)1/2
AU,
(3)
where L∗ and L⊙ are the luminosities of the cen-
tral star and the sun, respectively. In reality, disks
may be optically thick even after planetesimal for-
mation. However, we assume Equation (3) for sim-
plicity.
2.2. Fragmentation Outcome Model
Kobayashi & Tanaka (2010) showed erosive col-
lisions to dominate the collision cascade. We
should take into account such collisions prop-
erly. We assume that fragmentation outcomes are
scaled by the impact energy and hence the total
ejecta mass me produced by a single collision be-
tween m1 and m2 is given by a function of the di-
mensionless impact energy φ = m1m2v2/2(m1 +
m2)2Q∗
D, where v is the collisional velocity be-
tween m1 and m2 and Q∗
D is the specific en-
ergy needed for me = (m1 + m2)/2. Following
Kobayashi & Tanaka (2010) and Kobayashi et al.
(2010), we model
me
m1 + m2
=
φ
1 + φ
.
(4)
Inaba et al. (2003) derived me from the frag-
ment model developed by Wetherill & Stewart
3
D found by Benz & Asphaug
(1993) with a value of Q∗
(1999) for ice. Fig. 1 shows their model and Equa-
tion (4). As discussed in Kobayashi & Tanaka
(2010), most of the laboratory experiments and
the hydrodynamic numerical simulations of colli-
sional disruption showed me not to have a discon-
tinuity at φ = 1 (Housen et al. 1991; Takagi et al.
1984; Benz & Asphaug 1999). Therefore, Equa-
tion (4) includes erosive collisions (φ < 1) more
accurately.
1
0.1
0.01
)
2
m
+
1
m
(
/
e
m
0.001
0.01
0.1
1
10
φ
Fig. 1. -- The total ejecta mass me produced by
a single collision with m1 and m2 , as a function
of the dimensionless energy φ = m1m2v2/2(m1 +
m2)2Q∗
D. The solid line indicates Equation (4).
For reference, the dotted lines are shown for the
fragment model of Inaba et al. (2003) with m1 =
103 m2 = 4.2 × 1020 g.
The critical energy Q∗
D is given by
Q∗
D = Q0s r
1 cm!βs
+ Q0gρp r
1 cm!βg
+ Cgg
2Gm
r
, (5)
where r and m are the radius and mass of a body,
ρp is its density, and G is the gravitational con-
stant. The first term on the right-hand side of
Equation (5) is dominant for r . 104 -- 105 cm,
D of r . 107 cm,
the second term describes Q∗
and the third term controls Q∗
D for the larger
bodies. Benz & Asphaug (1999) performed the
hydrodynamical simulations of collisional disper-
sion for r = 1 -- 107 cm and provided the values of
Q0s, βs, Q0g, and βg. For r & 107 cm, Q∗
D is purely
4
determined by the gravitational binding energy,
being independent of material properties. The
collisional simulation for gravitational aggregates
yields Cgg ∼ 10 (Stewart & Leinhardt 2009).
2.3. Enhancement Radius by Atmosphere
Once a planetary embryo has grown larger than
the Moon, it acquires an atmosphere. It helps the
accretion of planetesimals or fragments onto an
embryo; small bodies are captured by the atmo-
sphere of the embryo.
Inaba & Ikoma (2003) provided an analytical
model for a density profile of the atmosphere. We
consider the atmosphere at a distance Re from
an embryo center. We assume that Re is much
smaller than that at the outer boundary of the at-
mosphere and that its temperature is much higher
than that at the boundary. The atmospheric den-
sity ρa is then proportional to R−3
(Mizuno 1980;
Stevenson 1982). Applying the temperature Tneb,
pressure Pneb, and density ρneb of the nebula in the
disk midplane as those at the outer boundary of
atmosphere, the density profile of the atmosphere
around an embryo with mass M is given by
e
ρa(Re)
ρneb
=
16πσSBGM T 4
neb
3κLePneb
4PnebRe(cid:19)3
(cid:18) GM ρneb
,
(6)
where κ is the opacity of the atmosphere and σSB
is the Stephan-Boltzmann constant. The plane-
tary luminosity Le mainly comes from the accre-
tion of bodies. We approximate
Le =
GM
R
dM
dt
,
(7)
where R is the embryo radius. To validate the as-
sumption of ρa ∝ R−3
, we will apply the complete
model by Inaba & Ikoma (2003) to our statistic
simulation and compare our analytical solutions
with the statistical simulations in Section 4.
e
When a body passes by a planetary embryo
with an atmosphere, the embryo can accrete the
body without direct collision due to the atmo-
sphere. The relative velocity between the body
and the embryo at infinity is determined by the
eccentricity e of the small body; it is given by evk
with the Keplerian velocity vk = pGM∗/a and
M∗ being the mass of a central star. The rela-
tive velocity is typically smaller than the surface
escape velocity of the embryo during the embryo
growth.
If the orbital energy of a body is suffi-
ciently reduced by the atmospheric gas drag, the
body is captured by the embryo. The maximum
radius r of bodies captured at distance Re is given
by (Inaba & Ikoma 2003)
r =
9ahM
6 + e2
ρa(Re)
ρp
,
(8)
where hM = (M/3M∗)1/3 is the reduced Hill
radius of the embryo and e = e/hM . Equa-
tion (8) is derived under the two-body approxima-
tion. Tanigawa & Ohtsuki (2010) confirmed that
Equation (8) is valid in the case where the three-
body effects are included.
Equation (8) means that Re is the effective col-
lisional radius of an embryo for bodies with radius
r. The enhanced radius of the embryo with atmo-
sphere is thus derived from Eqs. (6) -- (8) as
Re
R
=
F M 8/9
m1/9 M 1/3
,
(9)
where
F =(cid:20) π2aσSBT 4
nebG3
(e2 + 6)(3M∗)1/3κP 4
nebρ4
neb(cid:21)1/3
.
(10)
The enhancement factor Re/R given by Equa-
tion (9) is shown in Fig. 2, where the power-
law density profile given by Equation (6) is
compared with a more realistic profile given by
Inaba & Ikoma (2003). As we discuss later, plan-
etary embryos mainly grow through collisions with
planetesimals of the initial size or with fragments
of radius r ∼ 10 m. The enhancement factor cal-
culated with Equation (9) reproduces well the
more realistic one for km-sized or larger planetesi-
mals, but Equation (9) significantly overestimates
Re/R for fragments. However, since the accre-
tion rate due to collision with such fragments has
a weak dependence on the enhancement factor
(∝ (Re/R)1/2; see Equation (30)), this discrep-
ancy produces insignificant errors.
3. FINAL EMBRYO MASS
3.1.
Isolation Mass
Planetary embryos can grow until they have
accreted all planetesimals within their feeding
R
/
e
R
50
20
10
8
5
2
1
100
κ
=
1
c
m
2
/
g
κ
=
0
.
0
1
c
m
2
/
g
f = 0.0001
f
=
1
f
=
0
.
0
1
103
106
r [cm]
109
Fig. 2. -- The ratio of the enhanced radius of plan-
etary embryo to its physical radius with M = M⊕,
ρp = 1g cm−3, and M = 1 × 10−6M⊕/yr for
e = 4 in the MMSN disk around the star with
mass M⊙. The ratios are calculated by the formu-
lae of Inaba & Ikoma (2003) for the opacity ob-
tained from Equation (39) with the grain deple-
tion factor f = 10−4 -- 1 (solid lines) and by Equa-
tion (9) for the constant opacity κ = 0.01 cm2 g−1
and 1 cm2 g−1 (dotted lines).
zones. The width of a feeding zone is given
by the orbital separation of neighboring em-
bryos, b(2M/3M∗)1/3a, where b ≃ 10 is the
separation measured in their mutual Hill radii
(Kokubo & Ida 2000, 2002). The maximum mass
or "isolation mass" is Miso = 2πa2(2Miso/3M∗)1/3bΣs,0.
It can be expressed as
Miso = 2.8 b
× a
10!3/2 Σs,0
2.7 g cm−2!3/2
5 AU!3 M∗
M⊙!−1/2
M⊕, (11)
where M⊕ is the Earth mass and M⊙ is the
solar mass. The planetary embryo mass ap-
proaches the isolation mass if fragmentation is
ignored (Kokubo & Ida 2000, 2002). However, if
fragmentation is included, the embryo mass can
reach only about Mars mass for a MMSN disk
(Kobayashi et al. 2010).
5
3.2. Planetesimal Accretion
time τ is characterized as (Adachi et al. 1976)
As shown by Kobayashi et al. (2010), a plan-
etary embryo accretes planetesimals with masses
comparable to original ones or fragments resulting
from collisional grinding of planetesimals. In the
former case, a final embryo mass is determined by
the equilibrium between the accretion of planetes-
imals and their removal due to collisional grind-
ing. In the latter case, an embryo can grow un-
til fragments are depleted by the gas drag. Fol-
lowing Kobayashi et al. (2010), we want here to
derive final masses determined by the accretion
of planetesimals in the case with atmospheric en-
hancement, while we treat the fragment accretion
in Section 3.3.
At the oligarchic stage, embryos mainly grow
through collisions with planetesimals that domi-
nate the surface density. The growth rate of an
embryo with mass M is given by
dM
dt
= CaccΣsa2h2
M hPcoliΩk,
(12)
where ΩK is the Keplerian frequency and Cacc is
the correction factor on the order of unity. The
dimensionless collision rate hPcoli is formulated as
a function of the eccentricities e and inclinations
i of bodies accreted onto the embryo. We assume
e = 2i in this analysis.
Embryos have a constant ratio of their sepa-
rations to their Hill radii (Kokubo & Ida 1998).
When the ratio decreases as embryos grow, rela-
tively smaller embryos are culled and thereby re-
maining embryos keep the ratio constant. Sup-
posing the cull occurs instantaneously, the growth
rate of embryos due to the cull is estimated to be a
half of that from planetesimals. Therefore, we set
Cacc = 1.5 (e.g., Chambers 2006; Kobayashi et al.
2010).
For kilometer-sized or larger planetesimals,
their eccentricities e are controlled by the em-
The stirring rate
bryo stirring and gas drag.
is written as de2/dt = nM a2h4
M hPVSiΩk, where
nM is the surface number density of embryos and
the dimensionless stirring rate hPVSi is given by
hPVSi = CVSh2
M ln(Λ2 + 1)/e2 with CVS = 40 and
Λ = 5e3/96 for e ≫ hM (Ohtsuki et al. 2002).
Although ln(Λ2 + 1) in hPVSi is weakly dependent
on e, we adopt, in this analysis, ln(Λ2 + 1) ≃ 8,
with which value we can reproduce the formula of
Ohtsuki et al. (2002) for e = 3 -- 10. The gas-drag
τ =
2m
πr2CDρnebvk
,
(13)
where the dimensionless gas drag coefficient CD =
0.5 for km-sized or larger planetesimals. It should
be noted that τ is the stopping time due to gas
drag only when the relative velocity u between
gas and a body is equal to the Keplerian ve-
locity; hence τ is almost always much shorter
than the stopping time for realistic relative ve-
locities. The e-damping rate due to gas drag is
given by de2/dt = −Cgase3/τ with Cgas = 2.1
(Inaba et al. 2001). Using nM = (2πaδa)−1 with
the orbital separation of neighboring embryos of
δa = 21/3hM ab (Kokubo & Ida 2000) and equat-
ing the stirring and damping rates result in the
equilibrium eccentricity: 1
e =" CVS ln(Λ2 + 1)Ωkτ
24/3πbCgas
#1/5
.
(14)
Since we roughly estimate e ∼ (τ Ωk)1/5 from
Equation (14), the eccentricities of the kilometer-
sized and larger bodies are larger than hM , ac-
cording to the assumption e ≫ hM .
Taking into account the enhancement due
to the atmosphere,
the dimensionless colli-
sional probability for e ≫ hM is given by
(Greenzweig & Lissauer 1992; Inaba et al. 2001;
Inaba & Ikoma 2003)
hPcoli =
Ccol R
e2
Re
R
,
(15)
where Ccol = 36 and R = R/ahM = (9M∗/4πρp)1/3/a.
Inserting Eqs. (9) and (15) to Equation (12), we
obtain M as
dM
dt
= AcaM 7/6Σ3/4
s
,
(16)
where
Aca =" Cacca2Ccol RF Ωk
(3M∗)2/3e2m1/9 #3/4
.
(17)
1Ida & Makino (1993) and Thommes et al. (2003) pre-
sented a similar equation from the stirring timescale de-
rived by Ida & Makino (1993). We apply the formula of
Ohtsuki et al. (2002), which weakly depends on e through
ln(Λ2 + 1). However, since we adopt a constant value for
ln(Λ2 + 1) in this analysis, there is no substantial difference
between their and our treatment, except for the definition
of the coefficient for the viscous stirring.
6
As embryos grow, destructive collisions be-
tween planetesimals are induced by the stirring
of embryos and generate a lot of small fragments,
which produce further small bodies through mu-
tual collisions.
Since very small bodies re-
sulting from successive collisions are rapidly re-
moved by the gas drag, the collision cascade re-
duces the surface density of solids.
In the colli-
sion cascade, collisional fragmentation dominates
the mass flux along the mass coordinate. Since
the mass flux is independent of mass in a steady
state, the mass distribution of fragments follows a
power law and the power-law exponent α is given
D ∝ m−p
by α = (11 + 3p)/(6 + 3p) for e2/Q∗
(Kobayashi & Tanaka 2010). The steady-state
mass flux determines the surface density reduction
as (Kobayashi & Tanaka 2010; Kobayashi et al.
2010)
where M0 is the initial embryo mass. Note that
the derivation of Equation (21) assumed that the
planetesimal density reduction is caused by col-
lisional grinding, but the planetesimal accretion
onto embryos significantly contributes to the Σs-
reduction when the surface density of planetesi-
mals, Σs, is much smaller than that of embryos,
M nM . When an embryo reaches a final mass
Mca, Σs may be described as CΣs McanM with a
constant CΣs ≪ 1; hence
Σs
Σs,0
= CΣs(cid:18) Mca
Miso(cid:19)2/3
.
(22)
For CΣs ∼ 0.1, a final mass is consistent with sim-
ulations (Kobayashi et al. 2010). We thus set
CΣs = 0.1 to derive a final mass. From Eqs. (21)
and (22), we obtain a final embryo mass
dΣs
dt
= −BcaΣ2
s M 2(α−1)/3.
(2 − α)2Ωks123(α)
(18)
Mca =" 2(4α − 5)AcaC −1/4
Σs
3Bca
Σ−1/4
s,0 M 1/6
iso
#
3/2(α−1)
.
(23)
Bca =
where
s123(α) = Z ∞
0
m1/3
e2v2
k
2(3M∗)2/3Q∗
D(cid:19)α−1
,
(19)
×(cid:18)
(cid:20) φ
2 − b
φ−α
1 + φ
×
− φ ln
ǫφ
(1 + φ)2 + ln(1 + φ)(cid:21)
dφ,
(20)
p
and h0 = 1.1ρ−2/3
. For the derivation of Equa-
tion (18), we apply the fragmentation outcome
model of Kobayashi & Tanaka (2010);
ejecta
yielded by a single collision between m1 and m2
are characterised by their total mass me and their
power-law mass spectrum with an exponent b be-
low the mass mL = ǫ(m1 + m2)φ/(1 + φ)2, where
ǫ < 1 is a constant. The Σs reduction rate is in-
sensitive to ǫ and b (Kobayashi & Tanaka 2010).
We set b = 5/3 and ǫ = 0.2 in this paper.
Dividing Equation (16) by Equation (18) and
integrating, we obtain the relation between the
embryo mass M and the surface density Σs:
6
4α − 5hM (4α−5)/6 − M (4α−5)/6
0
i
= 4(Σ−1/4
s
− Σ−1/4
s,0
)
Aca
Bca
,
(21)
7
Here, we assume Mca ≫ M0.
For kilometer-sized or larger planetesimals,
D = Q0gρprβg with constants Q0g and βg. We
Q∗
apply Q0g = 2.1 erg cm3 g−2 and βg = 1.19 for ice
(Benz & Asphaug 1999) and e2 ≫ 6, and Equa-
tion (23) can then be re-written as
Mca = 1.8 × 10−2(cid:16) a
Q0g
5 AU(cid:17)2.8(cid:18)
2.1 erg cm3 g−2(cid:19)1.5(cid:18)
1.7 × 103 g cm−2(cid:19)1.41
fgasΣ1
×(cid:18)
×(cid:18)
m
4 × 1020 g(cid:19)0.63
0.01 g cm−2(cid:19)−0.51
κ
M⊕.
(24)
Since planetesimals grow before planetesimals'
fragmentation starts, planetesimal mass m is
slightly larger than initial planetesimal mass m0.
Kobayashi et al. (2010) showed that planetesimals
mainly accreting onto embryos have m = 100m0.
For m0 & 1023 g (r0 & 3 × 103 km), final embryo
masses exceed 10 M⊕ at 5 AU in a MMSN disk,
but embryos cannot reach it within a disk life-
time due to their slow growth. The final mass
Mca is independent of Σs,0, while high Σg,0 in-
creases Mca because gas drag highly damps e. For
Σ1 = 71 g cm−2 (10×MMSN), initial planetesi-
mals with r0 & 50 km can produce an embryo
with 10 M⊕ at 5 AU.
For comparison, we also show the final mass
Mc in the same situation but neglecting the atmo-
sphere (Kobayashi et al. 2010):
Mc = 0.10(cid:16) a
m
5 AU(cid:17)0.63(cid:18)
×(cid:18) ln(Σs,0/Σs)
×(cid:18)
4 × 1020 g(cid:19)0.48
(cid:19)1.21(cid:18)
1.7 × 103 g cm−2(cid:19)1.21
fgasΣ1
M⊕,
4.5
Q0g
2.1 erg cm3 g−2(cid:19)0.89
(25)
where ln(Σs,0/Σs) ≃ 4.5 is estimated from Equa-
tion(22) with CΣs = 0.1 for M = 0.1M⊕ in the
MMSN model. The collisional enhancement due
to the atmosphere is inefficient for m = 4 × 1020 g;
If m & 4 × 1022 g, the atmosphere
Mca < Ma.
contributes to embryo growth.
3.3. Fragment Accretion
As described above, planetesimals are ground
down by collision cascade and resulting small frag-
ments spiral into the central star by gas drag.
In the steady state of collision cascade, the sur-
face density of planetesimals is much larger than
that of fragments. However, when the grinding of
planetesimals is much quicker than the removal of
small fragments by gas drag, fragments accumu-
late at the low-mass end of collision cascade and
determine the total mass of bodies. Embryos then
grow through the accretion of such fragments.
The specific impact energy between equal-sized
bodies, e2v2
k/8, should be much smaller than Q∗
D
at the low-mass end; thus the typical fragments at
the low-mass end have
Keplerian velocity. The dimensionless viscous stir-
ring rate is given by hPVSi = hPVS,lowi = 73 for
e ≪ hM (Ohtsuki et al. 2002) and the damping
rate is expressed as de2/dt = −2ηe2/τ for e ≪ η
(Adachi et al. 1976). The equilibrium eccentricity
between stirring by embryos and damping by gas
drag is obtained as (Kobayashi et al. 2010)
e2 =
h3
M hPvs,lowiτ ΩK
27/3πbη
,
(27)
Using Eqs. (13), (26), and (27) under the Stokes
regime, we have the fragment mass mf at the low-
mass end of collision cascade:
mf = mf0M −3/2,
(28)
where
mf0 =" 221M∗bCLQ∗
D
hPVS,lowia2Ω3
K
c
lg,0 (cid:18) 3
4πρp(cid:19)1/3#3/2
.
(29)
For e ≪ hM, Ida & Nakazawa (1989) found
that the dimensionless collision rate for e ≪ hM
is given by hPcol,lowi = 11.3p R, where the coeffi-
cient is determined by Inaba et al. (2001). Since
the atmosphere effectively enhances an embryo ra-
dius for the accretion of bodies, the collision rate
is modified to be (Inaba & Ikoma 2003)
hPcoli = hPcol,lowir Re
R
.
(30)
We obtain the accretion rate of fragments by an
embryo,
M , from Eqs. (12) and (30) as
e2v2
k = CLQ∗
D,
(26)
dM
dt
= AfaM 43/42Σ6/7
s
,
where CL ∼ 1 is a constant. Although Kobayashi et al.
(2010) used CL = 1 to determine the typi-
cal fragment mass, we apply CL = 0.5 to cor-
rect a mistake of factor 2 in their e2.
Such
small fragments feel strong gas drag in Stokes
regime; CD = 5.5clg/ur, where c is the sound
velocity and lg = lg,0/ρg is the mean free path
of gas molecules with lg,0 = 1.7 × 10−9 g cm−2
(Adachi et al. 1976). The eccentricities of frag-
ments at the low-mass end are much smaller than
hM and η, where η = (vk − vgas)/vk is the de-
viation of the gas rotation velocity vgas from the
8
Afa = " F 1/2hPcol,lowiCacca2Ωk
(3M∗)2/3
m1/18
f0
(31)
. (32)
#6/7
Fragments with mf at the low-mass end of
collision cascade that dominate the surface den-
sity of solids Σs are no longer disrupted by col-
lisions and drift inward by gas drag. The drift
velocity is given by 2η2a/τ and then the Σs-
reduction rate due to the radial drift is expressed
as dΣs/dt = −2(9/4 − q)η2Σs/τ with the assump-
tion of Σs ∝ a−q. Since τ of fragments with mf is
determined by Equations (26) and (27), we have
(Kobayashi et al. 2010)
dΣs
dt
= −BfaΣsM,
Bfa = (cid:18) 9
4
− q(cid:19) hPvs,lowiΩkηv2
k
24/33πM∗CLbQ∗
D
(33)
. (34)
Since fragments are later produced by embryo
growth in an outer disk, the radial distribution
depends on time in contrast to the assumption of
Σs ∝ a−q. Nevertheless, the effect is negligible for
embryo growth unless the atmosphere is consid-
ered (Kobayashi et al. 2010). We discuss this ef-
fect with the atmospheric enhancement in §4 and
§5.
We can now obtain the final embryo mass Mfa
from M and Σs for fragment accretion, similar to
the case of planetesimal accretion. Integration of
Equation (31) divided by Equation (18) results in
Mfa =(cid:18) 41Afa
36Bfa(cid:19)42/41
Σ36/41
s,0
,
(35)
where we assume Mfa ≫ M0 and Σs,0 ≫ Σs. For
q = 3/2, we have
κ
0.01 g cm−3(cid:19)1/7
Mfa = 0.20(cid:16) a
×(cid:18) ficeΣ1
×(cid:18)
5 AU(cid:17)117/164(cid:18)
30 g cm−2(cid:19)36/41
3.1 × 106 erg g−1(cid:19)42/41
Q∗
D
M⊕.
(36)
Here, we adopted fice and Σ1 for the minimum
mass solar nebula model. The weak dependence
of Mfa on κ implies that the overestimate of Re/R
due to the power-law radial profile is insignificant,
as discussed in Section 2.3.
For the case without an atmosphere, Kobayashi et al.
(2010) derived a final mass for the fragment ac-
cretion,
Mf = 0.14(cid:16) a
×(cid:18)
30 g cm−2(cid:19)3/4
5 AU(cid:17)3/8(cid:18) ficeΣ1
3.1 × 106 erg g−1(cid:19)3/4
Q∗
D
M⊕.
(37)
Eqs.
(36) and (37) imply that the final masses
increase due to the atmosphere, but the enhance-
ment is insignificant; Mfa/Mf ≃ 1.4 -- 2 for 1 --
10×MMSN.
9
If we neglect the collisional enhancement due to
atmosphere, the final mass Mna is determined by
the larger of Mc and Mf (Kobayashi et al. 2010).
In the case with atmosphere, a final mass Ma is
also given by the larger of Mca and Mfa. The final
mass Ma is shown in Figs. 3 -- 5. For the initial
planetesimal radius r0 = 10 km, Ma is dominated
by Mfa inside the point where the line of Ma bends
in Fig. 3 and by Mca outside. The final mass Ma
is determined only by Mfa for r0 = 1 km (Fig. 4)
and by Mca for r0 = 100 km (Fig. 5) in the range
of interest.
]
M
[
s
s
a
M
o
y
r
b
m
E
100
10
1
0.1
0.01
100
10
1
0.1
0.01
100
10
1
0.1
0.01
Miso
Ma
Mna
Ma
Mna
Miso
Miso
Ma
Mna
4
7 10
30
Distance [AU]
3. -- Embryo masses with (circles) and
Fig.
without (squares) atmosphere after 107 years for
m0 = 4.2 × 1018 g (r0 = 10 km), as a func-
tion of distance form the central star. We set
Σ1 = 71 g cm−2 (top), Σ1 = 21 g cm−2 (middle),
and Σ1 = 7.1 g cm−2 (bottom). Solid lines in-
dicate Ma which is the larger of Mca and Mfa
for κ = 0.01cm2g−1. Dotted lines represent Mna
which is the larger of Mc and Mf . Thin lines show
Miso.
]
M
[
s
s
a
M
o
y
r
b
m
E
100
10
1
0.1
0.01
100
10
1
0.1
0.01
100
10
1
0.1
0.01
Ma
Ma
Mna
Miso
Mna
Miso
Miso
Ma
Mna
4
7 10
30
Distance [AU]
]
M
[
s
s
a
M
o
y
r
b
m
E
100
10
1
0.1
0.01
100
10
1
0.1
0.01
100
10
1
0.1
0.01
Miso
Mna
Ma
Ma
Miso
Mna
Miso
Ma
Mna
4
7 10
30
Distance [AU]
Fig. 4. -- Same as Fig. 3, but for m0 = 4.2×1015 g
(radii of 1 km).
Fig. 5. -- Same as Fig. 3, but for m0 = 4.2×1021 g
(radii of 100 km).
4. NUMERICAL SIMULATION
Regarding the method of numerical simulation,
we basically follow Kobayashi et al. (2010). The
method of Kobayashi et al. (2010) is briefly ex-
plained here. In the calculation, a disk is divided
into concentric annuli and each annulus contains
a set of mass batches. We set the mass ratio
between the adjacent batches to 1.2, which can
reproduce the collisional growth of bodies result-
ing from N -body simulation without fragmenta-
tion (Kobayashi et al. 2010) and the analytical so-
lution of mass depletion due to collisional grinding
(Kobayashi & Tanaka 2010). The mass and veloc-
ity evolution of bodies and their radial transport
are calculated as follows.
- The mass distribution of bodies evolves
through their mutual collisions that pro-
duce mergers and fragments. The total mass
of fragments ejected by a single collision is
10
given by Equation (4) and the remnant be-
comes a merger. The collision rates between
the bodies are calculated from the formulae
of Inaba et al. (2001).
- The random velocities given by e and i of the
bodies simultaneously evolve through their
mutual gravitational interactions, gas drag,
and collisional damping. The formulae of
Ohtsuki et al. (2002) are applied to describe
the changing rates of e and i. The gas-
drag damping rates of e and i are described
as functions of e, i, η, and τ according to
Inaba et al. (2001). To determine τ , we take
into account Stokes and Epstein drag as well
as a drag law with a quadratic dependence
on velocity. For the collisional damping,
both fragments and a merger resulting from
a single collision have the velocity dispersion
at the gravity center of colliding bodies.
- In each annulus there is a loss and gain
of bodies due to their inward drift. The
number loss rate from an annulus is given
by R (N (m)vdrift/∆a)dm, where vdrift is the
drift velocity of bodies, N (m)dm is the num-
ber of bodies with mass ranging from m
to m + dm in the annulus, and ∆a is the
width of the annulus. The bodies lost from
each annulus are added to the next inner
annulus. The drift velocity is given by
(Kobayashi et al. 2010)
vdrift =
2aη
τ
τ 2
stop
1 + τ 2
stop (cid:20) (2E + K)2
9π2
e2 +
4
π2 i2 + η2(cid:21)1/2
(38)
where E = 2.157, K = 1.211 and the dimen-
sionless stopping time τstop = Ωkτ /(e+i+η)
is adopted.
Σgas = Σgas,0 exp(−t/Tgas,dep), where Tgas,dep
is the gas depletion timescale, which we set
to Tgas,dep = 107 years. Assuming a constant
Σgas gives almost the same results for final em-
bryo masses, because we consider time spans
t ≤ Tgas,dep.
Fig. 6 shows the embryo-mass evolution at
6.4 AU for f = 0.01. Runaway growth initially oc-
curs; embryo mass exponentially grows with time
during the stage. The runaway-growth timescale
is proportional to r0/Σs,0 (Ormel et al. 2010a,b).
When the embryo masses exceed 0.001-0.01M⊕,
oligarchic growth starts. Since massive embryos
,
dynamically excite planetesimals, the reduction
of planetesimals due to collisional fragmentation
stalls the embryo growth (Kobayashi et al. 2010).
For Σ0 = 7.1 g cm−2 (MMSN), the fragmenta-
tion limits the final mass to about Mars mass (∼
0.1M⊕) and the atmosphere is insignificant. Once
embryo masses exceed the Mars mass, atmosphere
substantially accelerates the embryo growth. For
Σ0 ≥ 21 g cm−2 (3×MMSN), the atmosphere leads
to further embryo growth. Nevertheless, embryos
finally attain asymptotic masses.
Results for these simulations are summarised in
Fig. 3, where the embryo masses after 107 years are
compared to analytical formulae for final embryo
masses. Embryo masses finally reach Ma inside
5 AU (Σ0 = 7.1 g cm−2), 10 AU (Σ0 = 21 g cm−2),
and 20 AU (Σ0 = 71 g cm−2). However, embryos
exceed Ma inside 5 AU for Σ0 = 71 g cm−2. This
excess comes from the embryo growth through col-
lisional accretion with bodies drifting from out-
side, which effect we did not consider in the anal-
ysis described in Section 3. To confirm the contri-
bution from drifting bodies, we show the surface
density evolution in Fig. 7. For Σ0 = 71 g cm−2,
the surface density of solids increases after 2 ×
105 years. Since the drift timescale shortens in-
ward, bodies from outside cannot raise the surface
density unless embryos accrete them. Therefore,
the increase in the surface density implies that em-
bryo grows through the accretion of such bodies.
The initial mass m0 of planetesimals in the sim-
ulations depends on their formation process, which
is not well understood yet. We perform the em-
bryo growth starting from different m0 (Figs. 4
and 5). Small planetesimals are relatively easily
fragmented due to low Q∗
D and quickly ground
down to the low-mass end of collision cascade.
In this paper, we add a collisional enhance-
ment due to the atmosphere. Although the sim-
ple power-law radial density profile of the atmo-
sphere (Equation (6)) is used for the derivation of
final masses (Mca, Mfa), the simulation incorpo-
rates a more realistic profile provided by the for-
mulae of Inaba & Ikoma (2003). The opacity of
the embryo's atmosphere in their model is given
by κ = κgas + f κgr, where κgas is the gas opacity,
κgr is the opacity of grains having an interstellar
size distribution, and f is the grain depletion fac-
tor. Following Inaba & Ikoma, we adopt
0.01 + 4f cm2 g−1
0.01 + 2f cm2 g−1
0.01 cm2 g−1
for T ≤ 170 K,
for 170 K < T ≤ 1700 K,
for T > 1700 K.
κ =
(39)
The enhancement factor Re/R due to the atmo-
sphere is shown in Fig. 2.
We perform the simulations for embryo for-
mation starting from a monodisperse mass pop-
ulation of planetesimals of mass m0 and ra-
dius r0 with e = 2i = (2m0/M∗)1/3 and ρp =
1 g cm−3 around the central star of mass M⊙
with a set of eight concentric annuli at 3.2,
4.5, 6.4, 9.0, 13, 18, 25, and 35 AU contain-
ing Σgas and Σs
for q = 3/2. To compute
Q∗
D, we use Equation (5) with Q0s = 7.0 ×
107 erg g−1, βs = −0.45, Q0g = 2.1 erg cm3 g−2,
βg = 1.19, and Cgg = 9 (Benz & Asphaug 1999;
Stewart & Leinhardt 2009). We artificially ap-
ply the gas surface density evolution in the form
11
]
M
[
s
s
a
M
o
y
r
b
m
E
10
1
0.1
0.01
0.001
0.0001
10
1
0.1
0.01
0.001
0.0001
10
1
0.1
0.01
0.001
0.0001
104
105
106
107
Time [yr]
Fig. 6. -- Evolution of embryo mass at 6.4 AU
with m0 = 4.2 × 1018 g (r0 = 10 km) for
Σ0 = 71 g cm−2 (10×MMSN; top), 21 g cm−2
(3×MMSN; middle), and 7.1 g cm−2 (MMSN; bot-
tom). Solid lines show the case with atmosphere
and dotted lines represent the result without at-
mosphere.
The resulting fragments with low e actively ac-
For m0 = 4.2 × 1015 g
crete onto embryos.
(r0 = 1 km), embryos can reach a final mass Ma
in a relatively wide region inside 10 AU (MMSN),
20 AU (3×MMSN), and 30 AU (10×MMSN). On
the other hand,
large initial planetesimals de-
lay the runaway growth of embryos (Ormel et al.
2010a,b) and the following oligarchic growth is
also slower than that for small planetesimals be-
cause embryos mainly accrete original planetes-
imals rather than fragments with low e. For
r0 = 100 km, embryos attain the final masses
only inside 4 AU for 3×MMSN and inside 6 AU
for 10×MMSN, and embryos cannot reach final
masses beyond 2.7 AU in the MMSN disk. In ad-
dition, small bodies drifting from outside are ef-
fectively captured by embryos and thereby em-
12
]
2
m
c
/
g
[
Σs
80 Σ
50
1 = 71 g/cm2
20
10
8
5
2
1
0.8
0.5
102
Σ
1 = 21 g/cm2
Σ
1 = 7.1 g/cm2
103
105
104
Time [yr]
106
107
Fig. 7. -- The solid surface density evolution at
3.2 AU.
bryos exceed final masses Ma inside 4 AU for
10×MMSN.
In the case without an atmosphere,
initially
larger planetesimals can form massive embryos.
Since large planetesimals delay embryo growth,
embryos made from 100 km-sized initial planetesi-
mals can reach 10 M⊕ but the location is only in-
side 3 -- 4 AU even for 10×MMSN (Kobayashi et al.
2010). The case with the atmosphere shows a
similar dependence of the final embryo masses
on initial planetesimal mass. However, since the
atmosphere accelerates embryo growth, embryos
larger than 10 M⊕ are produced inside 8 -- 9 AU of
a 10×MMSN disk with 100 km-sized initial plan-
etesimals.
While the final masses of embryos exceed 10M⊕
for large initial planetesimals of r0 & 100 km, em-
bryos must reach the critical core mass within the
disk lifetime Tgas,dep to form gas giant planets.
The growth timescale is estimated to be M/ M ,
M is given by Equation (16). The crit-
where
ical distance ac inside which embryos can reach
10M⊕ is approximately obtained from the condi-
tion M/ M < Tgas,dep with M = 10M⊕,
ac = 9.6(cid:18) Tdep
×(cid:16) r0
107years(cid:19)20/39(cid:18)
100 km(cid:17)−1/3
AU,
Σ1
71 g cm−2(cid:19)23/39
(40)
where we adopt m = 100 m0 and q = 3/2.
For r0 = 100 km, the massive disk with Σ1 &
70 g cm−2 can form such large embryos around
10 AU. In addition, we estimate ac ∼ 5 AU
from Equation (40) for a 10×MMSN disk with
r0 = 103 km.
Indeed, the simulation with Σ1 =
71 g cm−2 and r0 = 100 km shows embryos cannot
reach 10M⊕ beyond 5 AU (see Fig. 8). Therefore,
the condition of 10×MMSN with r0 ∼ 100 km is
necessary to form gas giants around 10 AU.
]
M
[
s
s
a
M
o
y
r
b
m
E
100
10
1
0.1
0.01
Miso
Mna
4
7
10
30
Distance [AU]
Fig. 8. -- Same as Fig. 3, but for Σ1 = 71 g cm−2
with m0 = 4.2 × 1024 g (r0 = 1000 km). The final
mass Ma with atmosphere is estimated to be larger
than 200M⊕.
We also give a constraint on f . For f . 0.01, a
final mass is almost independent of f (see Fig. 9).
This is because the gas opacity dominates over the
grain opacity (see Equation (39)). For f = 1, em-
bryos at 3 -- 4 AU become larger due to the capture
of bodies drifting from outside, while final embryo
masses in the outer disk are similar to the case
without atmosphere. The condition of f . 0.01
is therefore necessary for gas giant formation in
the region 5 -- 10 AU and such low f is acceptable;
the depletion factor f should be much smaller
than unity after planetesimal formation.
In ad-
dition, a low-opacity atmosphere reduces the crit-
ical core mass (Mizuno 1980; Ikoma et al. 2000;
Hori & Ikoma 2010).
]
M
[
s
s
a
M
o
y
r
b
m
E
100
10
1
0.1
0.01
100
10
1
0.1
0.01
100
10
1
0.1
0.01
Ma
f = 0.0001
Miso
Mna
Miso
Ma
Mna
f = 0.01
Miso
Ma
Mna
f = 1
4
7 10
30
Distance [AU]
Fig. 9. -- The final embryo masses for f = 0.0001
(top), 0.01 (middle) and 1 (bottom), starting from
10×MMSN with m0 = 4.2 × 1021 g (r0 = 100 km).
Lines and symbols are the same as in Fig. 3, but
we apply κ = 1cm2 g−1 to derive Ma for f = 1.
5. DISCUSSION
We derived final embryo masses analytically
and numerically. They agree with each other quite
well in the inner disk where the embryo forma-
tion timescale is shorter than the nebula lifetime
(∼ 107 years). The analytical formula for final
masses Ma implies that initial planetesimal radii
should be larger than about 3×103 km to form em-
bryos with 10 M⊕ at 5 AU in a MMSN disk. How-
ever, the critical distance ac inside which embryos
reach 10 M⊕ within 107 years (Equation (40)) is
estimated to be much smaller than 5 AU; a massive
disk is likely to form gas giant planets. Embryos
inside 5 AU of a ∼ 10×MMSN disk exceed final
embryo masses Ma due to the accretion of small
bodies drifting from outside. In spite of such fur-
ther growth, embryos starting from small planetes-
13
imals cannot reach the critical core mass ∼ 10 M⊕.
In addition, further growth is insignificant beyond
5 AU. The formulae for Ma and ac suggest that
initial planetesimals with r0 ≃ 50 -- 700 km are nec-
essary for embryos to reach 10 M⊕ at 5 AU in the
10×MMSN disk.
Inaba et al. (2003) performed similar simula-
tions incorporating collisional fragmentation and
enhancement due to the embryo's atmosphere
and showed a planetary core with M > 10M⊕
could be produced around 5 AU with m0 =
4.2 × 1018 g (r0 = 10 km) for 10×MMSN. In
our simulation, embryos cannot reach 10M⊕ un-
der this condition and larger planetesimals are
necessary to form such massive embryos beyond
5 AU. As Kobayashi & Tanaka (2010) discussed,
Williams & Wetherill (1994) underestimated the
total ejecta mass produced by a single collision for
cratering; Inaba et al. adopted the fragmentation
model similar to theirs that Wetherill & Stewart
(1993) developed (see Fig. 1). Erosive collisions
shorten the depletion time of 10km-sized plan-
etesimals in collision cascade by a factor of 4 --
5 (Kobayashi & Tanaka 2010) and hence reduce
final embryo masses. As seen from Eqs. (24)
and (36), final embryo masses Mca, Mfa in-
crease with Q∗
D; the results of Inaba et al. cor-
respond to embryo masses for higher Q∗
D. Al-
applied Q∗
though we and Inaba et al.
D pro-
vided by Benz & Asphaug (1999), porous bodies
with r . 10 km may have much lower Q∗
D (e.g.,
Stewart & Leinhardt 2009; Machii & Nakamura
For initial planetesimals with radii &
2011).
100 km, Q∗
D of slightly larger bodies determines
final embryo masses and is almost entirely deter-
mined by the gravitational binding energy; the
uncertainty from their structure would be minor.
Therefore, such large planetesimals are possible to
produce cores for gas giant planets.
The mechanisms of planetesimal formation are
highly debated but, despite intensive effort, re-
main fairly unknown. The formation through col-
lisional coagulation in which dust smoothly grows
to planetesimals with r0 ∼ 1 km face barriers:
meter-sized objects should be lost to the central
star as a result of gas drag (Weidenschilling 1977;
Brauer et al. 2008), and further agglomeration of
cm-sized objects upon collision is problematic be-
cause of collisional bouncing (Guttler et al. 2010;
Zsom et al. 2010). Moreover, the electric repul-
sion may stop growth of smaller objects (Okuzumi
2009). A new scenario that allows one to over-
come the barriers has been proposed recently: self-
gravity of small particles accumulating in turbu-
lent structures of gaseous disks forms large plan-
etesimals of the order of 100 km (Johansen et al.
2007; Cuzzi et al. 2008). Not only do such large
planetesimals produce planetary cores exceeding
the critical core mass to form gas giant planets,
they may also be consistent with properties of mi-
nor bodies in the solar system. Indeed, the initial
planetesimals should be larger than 100 km to re-
produce the mass distribution of asteroids in the
main belt (Morbidelli et al. 2009).
For large planetesimals, a final embryo mass
given by Mca is large enough to start core ac-
cretion, while embryo growth is slow.
If the ra-
dial slope of surface density q = 3/2 like the
MMSN model, a massive disk with 10×MMSN
is necessary for embryos to reach the final mass
around 10 AU. However, observations of proto-
planetary disks infer their relatively flatter ra-
dial distributions over several hundred AU (e.g.,
Kitamura et al. 2002). In such a disk, dust grains
accumulate in an inner disk due to radial drift dur-
ing their growth, which increases the solid surface
density in the inner disk (Brauer et al. 2008). The
enhancement of solid surface density accelerates
embryo growth and hence embryos may achieve
the critical core mass in less massive disks.
To form gas giants via core accretion, rapid
gas accretion onto a core with ∼ 10 Earth masses
must occur prior to gas depletion. However, these
cores migrate inward due to their exchange of an-
gular momentum with the surrounding gas (Type
I). From linear analysis, the characteristic orbital
decay time of Earth-mass cores at several AU in
the MMSN model is about 1 Myr (Tanaka et al.
2002). Several processes to delay the timescale of
Type I migration have been pointed out, for exam-
ple, disk surface density transitions (Masset et al.
2006b), intrinsic turbulence (Nelson & Papaloizou
2004), and hydrodynamic feedback (Masset et al.
2006a). There is still uncertainty about this es-
timate of the migration time.
Indeed, the dis-
tribution consistent with observations of exoplan-
ets can be reproduced only if the timescale of the
type I migration is at least an order of magnitude
longer than that derived from the linear analysis
(Ida & Lin 2008). We should also investigate the
14
strength of such migration for the survival of cores
of gas giant planets in our future work.
6. SUMMARY
In this paper, we investigate the growth of plan-
etary embryos by taking into account, among oth-
ers, two effects that are of major importance. One
of them is collisional fragmentation of planetesi-
mals, which is induced by their gravitational inter-
action with planetary cores. Another effect is an
enhancement of collisional cross section of a grow-
ing embryo by a tenuous atmosphere of nebular
gas, which becomes substantial when an embryo
has reached about a Mars mass.
The main results are summarized as follows.
1. If the atmosphere is not taken into account,
collisional fragmentation suppresses plane-
tary embryo growth substantially. As a re-
sult, embryos cannot reach the critical core
mass of ∼ 10M⊕ needed to trigger rapid
gas accretion to form gas giants. The fi-
nal masses are about Mars mass in a MMSN
disk (Kobayashi et al. 2010). Embryo's at-
mosphere accelerates the embryo growth and
may increase the final embryo mass by up to
a factor of ten.
2. Planetary embryos attain their final masses
asymptotically. We have derived the final
mass analytically. The final mass of an em-
bryo is predicted to be the larger of Mca and
Mfa, which are given by Eqs. (24) and (36),
respectively. These final masses are in good
agreement with the results of statistical sim-
ulations.
3. Our solution indicates that an initial plan-
etesimal radius r0 & 3 × 103 km is necessary
to form a planetary core with 10 M⊕ at 5 AU
in a MMSN disk. However, such initially
large planetesimals delay embryo growth; a
massive disk is required to produce massive
cores within a disk lifetime. The analytical
solution for the final mass and the embryo
formation time show that planetesimals with
an initial radius of r0 ≃ 50 -- 700 km are likely
to produce such a large planetary core within
a disk lifetime at 5 AU for 10×MMSN.
4. The embryo growth depends on the disk
initial planetesimal sizes, and the
mass,
opacity of atmosphere. We have performed
statistical simulations to calculate the final
embryo masses over a broad range of param-
eters. We took the surface density of solids
at 1 AU in the range of Σ1 = 7.1 -- 71g cm−2
(1 -- 10×MMSN), initial planetesimal radius
r0 = 1 -- 1000 km, and the grain depletion
factor f in planetary atmosphere between
f = 10−4 -- 1. We found that planetary em-
bryos can exceed 10M⊕ within 8-9 AU for
10×MMSN, r0 = 100 km, and f ≤ 0.01.
Other sets of parameters cannot produce
massive cores at 5 -- 10 AU. For example, em-
bryo's mass can reach 6 M⊕ for r0 = 10 km
only inside 4 AU. Therefore, we conclude
that a massive disk (∼ 10×MMSN) with
r0 ∼ 100 km and f . 0.01 is necessary to
form gas giant planets around 5 -- 10 AU. This
condition for large embryo formation is in-
dependent of the material strength and/or
structure of bodies, because Q∗
D of 100km-
sized or larger bodies is largely determined
by their self-gravity.
We thank Chris Ormel for helpful discussions
and the reviewer, John Chambers, for useful com-
ments on the manuscript.
REFERENCES
Adachi, I., Hayashi, C., & Nakazawa, K. 1976,
Progress of Theoretical Physics, 56, 1756
Benz, W., & Asphaug, E. 1999, Icarus, 142, 5
Bodenheimer, P., & Pollack, J. B. 1986, Icarus,
67, 391
Bottke, W. F., Durda, D. D., Nesvorn´y, D.,
Jedicke, R., Morbidelli, A., Vokrouhlick´y, D.,
& Levison, H. F. 2005, Icarus, 179, 63
Brauer, F., Henning, T., & Dullemond, C. P. 2008,
A&A, 487, L1
Chambers, J. E. 2006, ApJ, 652, L133
15
Chambers, J. 2008, Icarus, 198, 256
Kenyon, S. J., & Bromley, B. C. 2009, ApJ, 690,
Cuzzi, J. N., Hogan, R. C., & Shariff, K. 2008,
ApJ, 687, 1432
Dohnanyi, J. S. 1969, J. Geophys. Res., 74, 2531
Fujiwara, A., Kamimoto, G., & Tsukamoto, A.
1977, Icarus, 31, 277
Greenzweig, Y., & Lissauer, J. J. 1992, Icarus, 100,
440
L140
Kitamura, Y., Momose, M., Yokogawa, S.,
Kawabe, R., Tamura, M., & Ida, S. 2002, ApJ,
581, 357
Kokubo, E., & Ida, S. 1996, Icarus, 123, 180
Kokubo, E., & Ida, S. 1998, Icarus, 131, 171
Kokubo, E., & Ida, S. 2000, Icarus, 143, 15
Guttler, C., Blum, J., Zsom, A., Ormel, C. W., &
Kokubo, E., & Ida, S. 2002, ApJ, 581, 666
Dullemond, C. P. 2010, A&A, 513, A56
Kobayashi, H., & Tanaka, H. 2010, Icarus, 206,
Hayashi, C. 1981, Progress of Theoretical Physics
735
Supplement, 70, 35
Kobayashi, H., Tanaka, H., Krivov, A. V., & In-
Holsapple, K. A. 1993, Annual Review of Earth
aba, S. 2010, Icarus, 209, 836
and Planetary Sciences, 21, 333
Levison, H. F., Thommes, E., & Duncan, M. J.
Hori, Y., & Ikoma, M. 2010, ApJ, 714, 1343
2010, AJ, 139, 1297
Housen, K. R., Schmidt, R. M., & Holsapple, K. A.
Machii, N., & Nakamura, A. M. 2011, Icarus, 211,
1991, Icarus, 94, 180
885
Ida, S., & Lin, D. N. C. 2008, ApJ, 673, 487
Masset, F. S., D'Angelo, G., & Kley, W. 2006,
Ida, S., & Makino, J. 1993, Icarus, 106, 210
Ida, S., & Nakazawa, K. 1989, A&A, 224, 303
Ikoma, M., Nakazawa, K., & Emori, H. 2000, ApJ,
537, 1013
Inaba, S., & Ikoma, M. 2003, A&A, 410, 711
Inaba, S., Wetherill, G. W., & Ikoma, M. 2003,
Icarus, 166, 46
Inaba, S., Tanaka, H., Nakazawa, K., Wetherill,
G. W., & Kokubo, E. 2001, Icarus, 149, 235
ApJ, 652, 730
Masset, F. S., Morbidelli, A., Crida, A., & Fer-
reira, J. 2006, ApJ, 642, 478
Mizuno, H. 1980, Progress of Theoretical Physics,
64, 544
Morbidelli, A., Bottke, W. F., Nesvorn´y, D., &
Levison, H. F. 2009, Icarus, 204, 558
Nelson, R. P., & Papaloizou, J. C. B. 2004, MN-
RAS, 350, 849
Ohtsuki, K. 1999, Icarus, 137, 152
Inaba, S., Tanaka, H., Ohtsuki, K., & Nakazawa,
K. 1999, Earth, Planets, and Space, 51, 205
Ohtsuki, K., Stewart, G. R., & Ida, S. 2002,
Icarus, 155, 436
Johansen, A., Oishi, J. S., Mac Low, M.-M.,
Klahr, H., Henning, T., & Youdin, A. 2007, Na-
ture, 448, 1022
Kenyon, S. J., & Bromley, B. C. 2004, AJ, 127,
513
Kenyon, S. J., & Bromley, B. C. 2008, ApJS, 179,
451
Okuzumi, S. 2009, ApJ, 698, 1122
Ormel, C. W., Dullemond, C. P., & Spaans, M.
2010a, ApJ, 714, L103
Ormel, C. W., Dullemond, C. P., & Spaans, M.
2010b, Icarus, 210, 507
Stevenson, D. J. 1982, Planet. Space Sci., 30, 755
Stewart, G. R., & Ida, S. 2000, Icarus, 143, 28
16
Stewart, S. T., & Leinhardt, Z. M. 2009, ApJ, 691,
L133
Takagi, Y., Mizutani, H., & Kawakami, S.-I. 1984,
Icarus, 59, 462
Tanaka, H., Takeuchi, T., & Ward, W. R. 2002,
ApJ, 565, 1257
Tanigawa, T., & Ohtsuki, K. 2010, Icarus, 205,
658
Thommes, E. W., Duncan, M. J., & Levison, H. F.
2003, Icarus, 161, 431
Weidenschilling, S. J. 1977, MNRAS, 180, 57
Weidenschilling, S. J. 2005, Space Sci. Rev., 116,
53
Weidenschilling, S. J. 2008, Physica Scripta Vol-
ume T, 130, 014021
Weidenschilling, S. J., Spaute, D., Davis, D. R.,
Marzari, F., & Ohtsuki, K. 1997, Icarus, 128,
429
Wetherill, G. W., & Stewart, G. R. 1989, Icarus,
77, 330
Wetherill, G. W., & Stewart, G. R. 1993, Icarus,
106, 190
Williams, D. R., & Wetherill, G. W. 1994, Icarus,
107, 117
Zsom, A., Ormel, C. W., Guttler, C., Blum, J., &
Dullemond, C. P. 2010, A&A, 513, A57
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
17
|
1003.4512 | 2 | 1003 | 2010-06-02T15:10:47 | The HAT-P-13 Exoplanetary System: Evidence for Spin-Orbit Alignment and a Third Companion | [
"astro-ph.EP"
] | We present new radial-velocity measurements of HAT-P-13, a star with two previously known companions: a transiting giant planet "b" with an orbital period of 3 days, and a more massive object "c" on a 1.2 yr, highly eccentric orbit. For this system, dynamical considerations would lead to constraints on planet b's interior structure, if it could be shown that the orbits are coplanar and apsidally locked. By modeling the Rossiter-McLaughlin effect, we show that planet b's orbital angular momentum vector and the stellar spin vector are well-aligned on the sky (lambda = -0.9 +/- 8.5 deg), suggesting that the planetary orbits are also well-aligned. The refined orbital solution favors a slightly eccentric orbit for planet b (e = 0.0142_{-0.0044}^{+0.0052}), although it is not clear whether it is apsidally locked with c's orbit (\Delta\omega = 48_{-38}^{+25} deg). We find a long-term trend in the star's radial velocity and interpret it as evidence for an additional body "d", which may be another planet or a low-mass star. The next inferior conjunction of c, when a transit may happen, is expected on JD 2,455,315.2 +/- 1.9 (centered on UT 17h 2010 April 28). | astro-ph.EP | astro-ph |
The HAT-P-13 Exoplanetary System:
Evidence for Spin-Orbit Alignment and a Third Companion
Joshua N. Winn1,2, John Asher Johnson3, Andrew W. Howard4,5, Geoffrey W. Marcy4,
G´asp´ar ´A. Bakos6,7, Joel Hartman6, Guillermo Torres6, Simon Albrecht1, Norio Narita2,8
ABSTRACT
We present new radial-velocity measurements of HAT-P-13, a star with two
previously known companions: a transiting giant planet "b" with an orbital
period of 3 days, and a more massive object "c" on a 1.2 yr, highly eccentric
orbit. For this system, dynamical considerations would lead to constraints on
planet b's interior structure, if it could be shown that the orbits are coplanar
and apsidally locked. By modeling the Rossiter-McLaughlin effect, we show that
planet b's orbital angular momentum vector and the stellar spin vector are well-
aligned on the sky (λ = 1.9±8.6 deg). The refined orbital solution favors a slightly
eccentric orbit for planet b (e = 0.0133 ±0.0041), although it is not clear whether
it is apsidally locked with c's orbit (∆ω = 36+27
−36 deg). We find a long-term trend
in the star's radial velocity and interpret it as evidence for an additional body
"d", which may be another planet or a low-mass star. Predictions are given for
the next few inferior conjunctions of c, when transits may happen.
Subject headings: planetary systems -- planetary systems: formation -- stars: in-
dividual (HAT-P-13) -- stars: rotation
1Department of Physics, and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute
of Technology, Cambridge, MA 02139
2Kavli Institute for Theoretical Physics, UCSB, Santa Barbara, CA 93106, USA
3Department of Astrophysics, California Institute of Technology, MC 249-17, Pasadena, CA 91125
4Department of Astronomy, University of California, Mail Code 3411, Berkeley, CA 94720
5Townes Postdoctoral Fellow, Space Sciences Laboratory, University of California, Berkeley, CA 94720
6Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138
7NSF Fellow
8National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan
-- 2 --
1.
Introduction
Precise radial-velocity measurements have revealed more than 30 multiple-planet sys-
tems (Wright 2010). However,
in only a few cases have transits been detected for any
of the planets in those systems. Those cases are potentially valuable because the transit
observables -- the times of conjunction, orbital inclination, and projected spin-orbit angle,
among others -- provide a much more complete description of a planetary system, which may
in turn give clues about its formation and evolution. The Corot-7 system has two orbiting
super-Earths, one of which transits (L´eger et al. 2009, Queloz et al. 2009). The HAT-P-7
system has a transiting hot Jupiter in a polar or retrograde orbit, as well as a longer-period
companion that could be a planet or a star (P´al et al. 2008, Winn et al. 2009, Narita et
al. 2010). The HAT-P-13 system, the subject of this paper, features a G4 dwarf star with
two previously known orbiting companions (Bakos et al. 2009). The inner companion (HAT-
P-13b, or simply "b" hereafter) is a transiting hot Jupiter in a 2.9 day orbit. The outer
companion ("c") has an eccentric 1.2 yr orbit and a minimum mass (Mc sin ic) of about 15
Jupiter masses, although its true mass (Mc) and orbital inclination (ic) are unknown. In
particular, transits of companion c have neither been observed nor ruled out.
Batygin, Bodenheimer, & Laughlin (2009) and Mardling (2010) showed that it may be
possible to use the observed state of the system to determine planet b's Love number k2, a
parameter that depends on the planet's interior density distribution. This would be of great
interest, as few other methods exist for investigating the interior structure of exoplanets.
The method is based on the theoretical expectation that tidal evolution has aligned the
apsides of the orbits of b and c. This method has not yet yielded meaningful constraints on
k2, partly because of the large uncertainty in the eccentricity of b's orbit. Another relevant
parameter is the mutual inclination between the orbits, which is not known at all.
Radial-velocity observations are usually powerless to determine mutual inclinations,
unless the planets are in a mean-motion resonance (see, e.g., Correia et al. 2010). However,
for a transiting planet it is possible to assess the alignment between the orbit and the stellar
equator through the Rossiter-McLaughlin (RM) effect. A system with mutually inclined
planetary orbits might also be expected to have large angles between the orbits and the
stellar equator. In particular, Mardling (2010) presented a formation scenario for HAT-P-13
involving gravitational scattering by a putative third companion, which could have caused
large mutual inclinations and a large stellar obliquity.
In this paper we present new radial-velocity measurements of HAT-P-13 bearing on all
these issues. The new data are presented in § 2. Our analysis is presented § 3, and includes
evidence for a third companion "d" (§ 3.1), refined estimates of the eccentricity and apsidal
orientation of b's orbit (§ 3.2), modeling of the RM effect (§ 3.3), and updated predictions
-- 3 --
for the next inferior conjunction (possible transit window) of companion c (§ 3.4). In § 4 we
discuss the implications for further dynamical investigations of HAT-P-13.
2. Observations
We observed HAT-P-13 with the High Resolution Spectrograph (HIRES; Vogt et al. 1994)
on the Keck I 10m telescope, using the same instrument settings and observing protocols that
were used by Bakos et al. (2009) and that are used by the California Planet Search (Howard
et al. 2009). In particular, we used the iodine gas absorption cell to calibrate the instrumen-
tal point-spread function and the wavelength scale. The total number of new spectra is 75,
which are added to the 30 spectra presented by Bakos et al. (2009). Of the new spectra, 40
were obtained on the night of 2009 Dec 27-28, spanning a transit of HAT-P-13b, and were
gathered to measure the RM effect. The other 35 were obtained on arbitrary nights. They
extend the timespan of the data set by approximately 1 yr, and thereby help to refine the
orbital parameters.
The radial velocity (RV) of each spectrum was measured with respect to an iodine-free
template spectrum, using the algorithm of Butler et al. (1996) with subsequent improve-
ments. All of the spectra obtained by Bakos et al. (2009) were re-reduced, for consistency.
Measurement errors were estimated from the scatter in the fits to individual spectral seg-
ments spanning 2 A. The RVs are given in Table 1, and plotted as a function of time in
Figures 1-2. The model curves appearing in those figures are explained in § 3.
Our model for the radial-velocity data took the form
3. Analysis
Vcalc(t) = Vb(t) + Vc(t) + VRM(t) + γ + γ(t − t0),
(1)
where Vcalc is the calculated RV, Vb and Vc are the radial velocities of non-interacting Keple-
rian orbits, VRM is the transit-specific "anomalous velocity" due to the Rossiter-McLaughlin
effect (§ 3.3), and {γ, γ, t0} are constants. The first constant, γ, specifies the velocity off-
set between the system barycenter and the arbitrary template spectrum that was used to
calculate RVs. The second constant,
γ, allows for a constant radial acceleration, and was
included because models with γ = 0 did not fit the data (§ 3.1). We interpret γ as the
acceleration produced by a newly-discovered long-period companion "d". The third con-
stant, t0, is an arbitrary reference time that was taken to be the time of the first RV datum
(BJD 2,454,548.80650).
-- 4 --
Fig. 1. -- Radial velocity variation of HAT-P-13. Top. -- Measured RVs, and the best-
fitting model. The model consisted of two Keplerian orbits and did not allow for any addi-
tional acceleration ( γ ≡ 0). Bottom. -- Residuals. The poorness of the fit, and the pattern of
residuals, are evidence for a third companion.
Our RM model was based on that of Winn et al. (2005), in which simulated spectra
are used to calibrate the relation between the phase of the transit and the measured radial
velocity. For this case we used the relation
∆V (t) = −(v sin i⋆) δ(t)"0.9833 − 0.0356(cid:18) vp(t)
v sin i⋆(cid:19)2# ,
(2)
where v sin i⋆ is the sky-projected stellar rotation speed, δ is the fractional loss of light, and
vp is the radial velocity of the portion of the stellar photosphere that is hidden by the planet.
To calculate vp we assumed that the stellar photosphere rotates uniformly with an angle λ
between the sky projections of the spin vector and the orbital angular momentum vector
(see, e.g., Ohta et al. 2005 or Gaudi & Winn 2007).
Since δ depends on the planet-to-star radius ratio Rp/R⋆, orbital inclination i, and
impact parameter btra, all of which are more tightly bounded by observations of photometric
-- 5 --
Fig. 2. -- Radial velocity variation of HAT-P-13. Top. -- Measured RVs, and the best-
fitting model, this time allowing for a constant radial acceleration ( γ) in addition to two
Keplerian orbits. The best fitting value of γ was 17.5 m s−1 yr−1. Bottom. -- Residuals.
transits than by the RM effect, we simultaneously fitted a composite i′-band transit light
curve based on the photometric data of Bakos et al. (2009). For the photometric model we
assumed a quadratic limb-darkening law and used the analytic formula of Mandel & Agol
(2002), as implemented by P´al (2008). Because the photometric data are not precise enough
to constrain both of the limb-darkening coefficients u1 and u2, we fixed u2 ≡ 0.3251, the
value obtained by interpolating the Claret (2004) tables, and allowed u1 to vary freely.1 For
the RM model, we used a linear law with a fixed coefficient of 0.72, as appropriate for the
V -band, the approximate spectral range from which the RV signal is drawn.
All together there were 18 adjustable parameters, of which 12 were controlled almost
exclusively by the RV data, and 6 by the photometric data. The data set had 105 RVs and
107 flux data points. Thus, the total number of degrees of freedom was 194, of which 93
1The result, u1 = 0.269 ± 0.076, was consistent with the tabulated value of 0.3068.
-- 6 --
pertained to RVs and 101 to photometry.
We determined the best-fitting parameter values and their 68.3% confidence limits with
a Monte Carlo Markov Chain algorithm that we have described elsewhere (see, e.g., Winn
et al. 2007). Uniform priors were adopted for all parameters except for b's time of transit
and orbital period, for which we adopted Gaussian priors based on the ephemeris of Bakos
et al. (2009). We doubled the quoted errors in the ephemeris, out of concern that systematic
errors or transit-timing variations have affected the results. The likelihood was taken to be
exp(−χ2/2) with
χ2 =
105
Xi=1 (cid:20) vi(obs) − vi(calc)
σi
(cid:21)2
+
107
Xj=1(cid:20) fj(obs) − fj(calc)
σj
(cid:21)2
= χ2
v + χ2
f ,
(3)
where vi(obs) and σi are the RV data and associated uncertainties, vi(calc) are the calculated
RVs; fj(obs) and σj are the flux data and associated uncertainties; and fj(calc) are the
calculated fluxes.
Each flux uncertainty σi was taken to be the scatter in the ≈16 data points contributing
to each 4 min time bin. Each RV uncertainty σj was taken to be the quadrature sum of the
internally-estimated measurement error and a "jitter" term of 3.4 m s−1. The jitter term was
set by the requirement χ2
v = 93, the relevant number of degrees of freedom, and is consistent
with the empirical jitter estimates of Wright (2005) for stars similar to HAT-P-13. In the
best-fitting model, χ2
f = 109.5 and χ2
v = 93.0, the rms photometric residual was 470 ppm
and the rms RV residual was 3.6 m s−1.
Table 2 gives the results for the parameter values. Figure 3 shows the RVs as a function
of the orbital phases of b and c, expressed in days. In the left panel, the RVs are plotted
as a function of the orbital phase of b, after subtracting the calculated contributions to the
RV from companions c and d.
(The contribution due to d is a linear function of time.)
Likewise, the right panel of Figure 3 shows the RVs as a function of the orbital phase of c,
after subtracting the calculated contributions from b and d. Figure 4 shows the data over
a restricted time range centered on b's transit. The top panel shows the light curve. The
bottom panel shows the data after subtracting the calculated orbital RV, thereby isolating
the RM effect.
3.1. Evidence for a third companion
The extra acceleration, γ, was included in the RV model because a model consisting of
only two Keplerian orbits gave an unacceptable fit to the data. With γ = 0, the RV-specific
-- 7 --
Fig. 3. -- Radial-velocity variation as a function of orbital phase. Left. -- RV variation
as a function of the orbital phase of planet b after subtracting the calculated variation due
to c and d. Right. -- RV variation as a function of the orbital phase of c, after subtracting
the calculated variation due to b and d.
portions of the data and model had χ2
v/Ndof,v =
4.9). The pattern of residuals is displayed in Figure 1. There are large and time-correlated
residuals that are not easily attributed to stellar jitter or underestimated measurement errors.
v = 458.6 and 94 degrees of freedom (χ2
In contrast, when γ was allowed to vary freely, the best-fitting model had γ = 17.5 m s−1 yr−1,
v = 93 with 93 degrees of freedom. The exact match between χ2
and χ2
v and Ndof,v is not sig-
nificant in itself, as it follows from our choice of 3.4 m s−1 for the jitter term. However, it
is significant that an acceptable fit was found for a choice of jitter term that is in line with
observations of similar stars. Even more significant is that the correlated pattern of residuals
vanished. As shown in Figure 2, the residuals scattered randomly around zero.
The failure of the two-Keplerian model, and the success of a model with an additional
radial acceleration, is evidence for a third companion to HAT-P-13 ("d") with a long orbital
period. With the limited information available, though, the properties of d are largely
unknown. Assuming its orbit to be nearly circular, and its mass to be much smaller than
that of the star, we may set γ ∼ GMd sin id/a2
d to give an order-of-magnitude constraint
(cid:18)Md sin id
MJup (cid:19)(cid:16) ad
10 AU(cid:17)−2
∼ 9.8,
(4)
where ad is the orbital distance. By this standard, the newly discovered object could be a
2.5 MJup planet at 5 AU, or a 10 MJup planet at 10 AU, or a 90 MJup (0.09 M⊙) star at
30 AU, etc. The properties of d could be substantially different depending on its eccentricity,
argument of pericenter, and time of conjunction. Orbits closer than ∼5 AU would be subject
-- 8 --
Fig. 4. -- Transit photometry and radial-velocity variation. Top. -- A composite
transit light curve based on the i′-band photometric data of Bakos et al. (2009). Also
plotted are the best-fitting model and the residuals. Bottom. -- The apparent RV variation
observed during the transit phase, after subtracting the calculated contributions due to
orbital motion. The observed variation is interpreted as the anomalous velocity due to the
Rossiter-McLaughlin effect.
-- 9 --
to additional constraints by the requirement of dynamical stability.
More information about d could be gleaned from any significant curvature in the RV
signal, beyond the effects of the two Keplerian orbits and a linear trend. We experimented
with models that include a "jerk" parameter, γ, finding this parameter to be highly covariant
with the mass, orbital period, and eccentricity of c. More elaborate models and detailed
constraints on companion d will only be justified after another few years of observing, when
the properties of companion c will have been well established. The uncertainties given in
Table 2 must therefore be understood as subject to the assumption that d is producing no
significant RV curvature.
3.2. Orbital eccentricities
Figure 5 shows the results for the orbital eccentricities. Planet c's orbit is strongly
eccentric, with ec = 0.6616 ± 0.0054. Planet b's orbit is nearly circular, with eb = 0.0133 ±
0.0041. To assess the significance of the detection of eccentricity, it is simpler to examine
the components of the eccentricity vector because they obey Gaussian distributions, while
e obeys a Rayleigh distribution (see, e.g., Shen & Turner 2008). We found eb cos ωb =
−0.0099 ± 0.0036 (i.e., nonzero with 2.8σ significance) and ec sin ωc = −0.0060 ± 0.0069. The
eccentricity of b's orbit is right on the edge of detectability.2
Because of the low significance of this detection, it is impossible to draw a firm conclusion
about whether its orbit is aligned with that of companion c. Our result is ∆ω ≡ ωb − ωc =
36+27
−36 degrees. The red dashed lines in Figure 5 show the 3σ allowed region for ωc. The lines
intersect the allowed region for planet b, as shown in the right panel. Most of the uncertainty
in ∆ω arises from the poorly determined orientation of b's orbit.3
The best way to check on the eccentricity of b's orbit would be to observe an occultation
with the Spitzer Space Telescope. The timing of occultations would allow eb cos ωb to be de-
termined with an accuracy of about 0.001, several times better than the RV result. However,
even after such an observation, considerable uncertainty would remain in ∆ω because the
accuracy in eb sin ωb would not be much improved.
2For this reason the results are also sensitive to the choice of priors for the fitting parameters. The results
If
described in this section and given in Table 2 are based on uniform priors for eb cos ωb and eb sin ωb.
instead uniform priors are adopted for eb and ωb, then we find eb = 0.0119 ± 0.0040.
3If we assume the orbits are apsidally locked, and repeat the fitting procedure with the requirement
∆ω = 0, we find eb = 0.0104 ± 0.0032.
-- 10 --
Fig. 5. -- Results for the orbital eccentricities. The middle panel displays the results
for b and c, while the left and right panels zoom in on the results each object. The contours
enclose 68%, 95%, and 99.73% of the MCMC samples. The dashed lines show the 99.73%
confidence range for the apsidal orientation of c's orbit; they allow a visual assessment of
the degree of apsidal alignment, and show that the limiting uncertainty in ∆ω is the large
uncertainty in eb sin ωb.
3.3. The Rossiter-McLaughlin effect
The RV data obtained during transits exhibit a prograde Rossiter-McLaughlin effect:
an anomalous redshift for the first half of the transit, followed by an anomalous blueshift for
the second half. The fit to the data is shown in Figure 4, and the resulting constraints on λ
and v sin i⋆ are shown in Figure 6. The finding of λ = 1.9 ± 8.6 deg implies a close alignment
between the rotational angular momentum of the star, and the orbital angular momentum
of the planet, at least as projected on the sky. Our result for the projected stellar rotation
velocity, v sin i⋆ = 1.66 ± 0.37 km s−1, is about 1σ smaller than the result of 2.9 ± 1.0 km s−1
reported by Bakos et al. (2009).
3.4.
Inferior conjunction of planet c
It is not yet known whether the inclination of c's orbit is close enough to 90◦ for transits
to occur. Observations of transits would reveal the mass and radius of the companion, allow
a more precise characterization of its orbit, and place constraints on the mutual inclination
of orbits b and c.
Using our results we predicted the times of inferior conjunctions of planet c, which is
when transits would occur. The accuracy of the predicted time is limited by correlations
with the uncertainties in c's velocity semiamplitude and eccentricity (see Figure 7). Table 3
-- 11 --
Fig. 6. -- Results for the Rossiter-McLaughlin parameters, based on our MCMC
analysis of the RV data. The contours represent 68%, 95%, and 99.73% confidence limits,
and the one-dimensional (marginalized) posterior probability distributions are shown on the
sides of the contour plot.
gives the results. The quoted uncertainties represent 1σ (68.3%) confidence levels. It would
be prudent to keep the star under continuous photometric surveillance for the entire 3σ time
range, at least. The maximum transit duration is 14.9 hr.
4. Discussion
HAT-P-13 was already a noteworthy system, as the first known case of a star with a
transiting planet and a second close companion. We have presented evidence for a third
companion in the form of a long-term radial acceleration of the star. The properties of the
newly discovered long-period companion will remain poorly known until additional RV data
are gathered over a significant fraction of its orbital period. Our analysis of the Rossiter-
McLaughlin effect shows that planet b's orbital axis is aligned with the stellar rotation axis,
-- 12 --
Fig. 7. -- Results for the timing of the inferior conjunction of companion c, based
on our MCMC analysis of the RV data. The top panel shows the one-dimensional (marginal-
ized) posterior probability distribution. The lower two panels illustrate the correlation with
the other poorly-determined parameters of companion c. The contours represent 68%, 95%,
and 99.73% confidence limits.
as projected on the sky. Our new data also agree with the previous finding that the orbit of
planet b is slightly eccentric.
The latter two findings are relevant to the second reason why HAT-P-13 is noteworthy:
its orbital configuration may represent an example of two-planet tidal evolution. In this sce-
nario, first envisioned by Wu & Goldreich (2002) and investigated further by Mardling (2007),
tidal circularization of the inner planet's orbit is delayed due to gravitational interactions
with the outer planet. The interactions drive the system into a state of apsidal alignment,
where it remains as both orbits are slowly circularized. As it turned out, the specific plane-
tary system that inspired Wu & Goldreich (2002) was irrelevant to their theory, because the
"outer planet" was found to be a spurious detection (Butler et al. 2002).
Batygin et al. (2009) welcomed HAT-P-13 as a genuine system that followed the path
predicted by Wu & Goldreich (2002), and with the additional virtue that the inner planet is
transiting. For this interpretation to be valid, the apsides of b and c must be aligned, whereas
-- 13 --
we have found the angle between the apsides to be 36+27
−36 deg, differing from zero by 1σ. We
do not consider this result to be significant enough to draw a firm conclusion, especially in
light of the uncertainties due to the ad hoc stellar jitter term and our simplified treatment of
the influence of companion d. Further RV monitoring and observations of occultations are
needed to make progress.
Batygin et al. (2009) also showed that the existence of transits would allow for an
empirical estimate of the tidal Love number k2 of planet b, as mentioned in § 1. The
requirement that the apsidal precession rates of b and c are equal leads to a condition on k2,
because b's precession rate is significantly affected by its tidal bulge. Subsequent work by
Mardling (2010) showed that for a unique determination of k2 it is necessary for the mutual
inclination ∆i between orbits b and c to be small. If instead the orbits are mutually inclined,
then tidal evolution drives the system into a state in which eb and ∆i undergo oscillations:
a cycle in parameter space, instead of a fixed point. Furthermore, Mardling (2010) argued
that a large mutual inclination should be considered plausible, or even likely, given c's high
eccentricity. She proposed that b and c began with nearly circular and coplanar orbits, but
c's orbit was strongly perturbed by an interaction with a hypothetical outer planet. Those
same perturbations would likely have tilted c's orbit.
The relation, if any, between the newly-discovered HAT-P-13d and Mardling's hypo-
thetical outer planet is unclear. In her scenario, the outer planet is ejected from the system.
This seems important to the scenario, as otherwise d would continue interacting with c, and
interfere with the tidal evolution of b and c. Thus, unless d's pericenter was somehow raised
to avoid further encounters with c, it does not seem likely to have played the role envisioned
by Mardling (2010). Of course the scenario could still be correct even if the third companion
d was not the scattering agent; a fourth (ejected) companion may have been responsible.
Our study of the Rossiter-McLaughlin effect pertains to the angle ψ⋆,b between planet b's
orbit and the stellar equator, and has no direct bearing on the angle ∆i between the orbital
planes of b and c. However, there is an indirect connection, through the nodal precession
that would be caused by mutually inclined orbits. As shown by Mardling (2010), planet
b is far enough from the star that its orbital precession rate is likely to be dominated by
the torque from c, rather than the quadrupole moment J2 of the star. The critical orbital
distance inside which the stellar torque is dominant is ∼(2J2a2
cMc/M⋆)0.2 (Burns 1986),
which is 0.020 AU assuming J2 = 2 × 10−7 as for the Sun. This is smaller than the actual
orbital distance of 0.043 AU. Hence if ∆i were large, then b's orbit would nodally precess
around c's orbital axis, which would cause periodic variations in ψ⋆,b. Therefore, at any given
moment in the system's history, we would be unlikely to observe a small value of ψ⋆,b unless
∆i were small. However, it is impossible to draw firm conclusions about ∆i because of the
-- 14 --
dependence on initial conditions, the possible effects of companion d, and the fact that only
the sky-projected angle λ is measured rather than the true obliquity ψ⋆,b.
It may be possible to estimate ∆i based on transit timing variations of planet b (Nesvorn´y
& Beaug´e 2010; Bakos et al. 2009). An even more direct estimate of ∆i could be achieved
if transits of c were detected. The existence of transits would show that ic is nearly 90◦, as
is ib. This would suggest ∆i is small, although it would still be possible that the orbits are
misaligned and their line of nodes happens to lie along the line of sight. The most defini-
tive result would be obtained by observing the Rossiter-McLaughlin effect during transits
of c, and comparing c's value of λ with that of planet b. In effect, the rotation axis of the
star would be used as a reference line from which the orientation of each orbit is measured
(Fabrycky 2009). This gives additional motivation to observe HAT-P-13 throughout the
upcoming conjunctions of companion c.
We thank Dan Fabrycky for helpful conversations, especially about the dynamical im-
plications of our results. We also thank Debra Fischer and John Brewer for investigating
the spectroscopic determination of the stellar rotation rate. We are grateful to Scott Gaudi
and Greg Laughlin for comments on the manuscript.
J.N.W. gratefully acknowledges support from the NASA Origins program through award
NNX09AD36G and the MIT Class of 1942. A.W.H. acknowledges a Townes Postdoctoral
Fellowship from the Space Sciences Laboratory at UC Berkeley. G.A.B. was supported by
NASA grant NNX08AF23G and an NSF Astronomy & Astrophysics Postdoctoral Fellowship
(AST-0702843). G.T. acknowledges partial support from NASA grant NNX09AF59G. S.A.
acknowledges the support of the Netherlands Organisation for Scientific Research (NWO).
N.N. was supported by a Japan Society for Promotion of Science (JSPS) Fellowship for
Research (PD: 20-8141). J.N.W. and N.N. were also supported in part by the National
Science Foundation under Grant No. NSF PHY05-51164 (KITP program "The Theory and
Observation of Exoplanets" at UCSB).
The data presented herein were obtained at the W.M. Keck Observatory, which is oper-
ated as a scientific partnership among the California Institute of Technology, the University
of California, and the National Aeronautics and Space Administration, and was made pos-
sible by the generous financial support of the W.M. Keck Foundation. We extend special
thanks to those of Hawaiian ancestry on whose sacred mountain of Mauna Kea we are priv-
ileged to be guests. Without their generous hospitality, the Keck observations presented
herein would not have been possible.
Facilities: Keck:I (HIRES)
-- 15 --
REFERENCES
Bakos, G. ´A., et al. 2009, ApJ, 707, 446
Batygin, K., Bodenheimer, P., & Laughlin, G. 2009, ApJ, 704, L49
Burns, J. 1986, in Satellites, eds. J. Burns and M. Matthews (University of Arizona Press),
p. 117
Butler, R. P., Marcy, G. W., Williams, E., McCarthy, C., Dosanjh, P., & Vogt, S. S. 1996,
PASP, 108, 500
Butler, R. P., et al. 2002, ApJ, 578, 565
Claret, A. 2004, A&A, 428, 1001
Correia, A. C. M., et al. 2010, A&A, 511, A21
Fabrycky, D. C. 2009, in Transiting Planets, eds. F. Pont, D. Sasselov, M. Holman, Proc.
IAU Symposium 253, p. 173-179 (Cambridge Univ. Press)
Gaudi, B. S., & Winn, J. N. 2007, ApJ, 655, 550
Howard, A. W., et al. 2009, ApJ, 696, 75
L´eger, A., et al. 2009, A&A, 506, 287
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Mardling, R. A. 2007, MNRAS, 382, 1768
Mardling, R. A. 2010, MNRAS, in press [arXiv:1001.4079]
Narita, N., et al. 2010, PASJ, in press [arXiv:1004.2458]
Nesvorn´y, D., & Beaug´e, C. 2010, ApJ, 709, L44
Ohta, Y., Taruya, A., & Suto, Y. 2005, ApJ, 622, 1118
P´al, A. 2008, MNRAS, 390, 281
P´al, A., et al. 2008, ApJ, 680, 1450
Queloz, D., et al. 2009, A&A, 506, 303
Shen, Y., & Turner, E. L. 2008, ApJ, 685, 553
-- 16 --
Vogt, S. S., et al. 1994, Proc. SPIE, 2198, 362
Winn, J. N., et al. 2005, ApJ, 631, 1215
Winn, J. N., Holman, M. J., & Fuentes, C. I. 2007, AJ, 133, 11
Winn, J. N., Johnson, J. A., Albrecht, S., Howard, A. W., Marcy, G. W., Crossfield, I. J.,
& Holman, M. J. 2009, ApJ, 703, L99
Wright, J. T. 2005, PASP, 117, 657
Wright, J. T. 2010, EAS Publications Series, 42, 3
Wu, Y., & Goldreich, P. 2002, ApJ, 564, 1024
This preprint was prepared with the AAS LATEX macros v5.2.
-- 17 --
Table 1. Relative Radial Velocity Measurements of HAT-P-13
BJD
RV [m s−1]
Error [m s−1]
2454548.80650
2454548.90850
2454602.73396
2454602.84691
2454603.73415
2454603.84324
2454633.77241
2454634.75907
2454635.75475
2454636.74969
2454727.13850
2454728.13189
2454778.07301
2454779.08373
2454780.09368
2454791.11129
2454809.99575
2454839.06085
2454865.02660
2454867.90311
2454928.83635
2454955.86964
2454956.86327
2454963.85163
2454983.74976
2454984.76460
2454985.73856
2454986.76358
2454988.74066
2455109.11745
2455110.10818
2455134.11719
2455135.13125
2455164.01155
2455172.12118
2455173.02454
2455188.04447
2455189.08587
2455189.98539
2455191.11450
2455192.02847
2455193.85943
2455193.86475
2455193.86961
2455193.94390
2455193.94850
2455193.95323
2455193.95775
2455193.96234
2455193.96702
2455193.97167
2455193.97628
2455193.98097
2455193.98536
2455193.98980
2455193.99433
2455193.99888
2455194.00354
2455194.00825
2455194.01271
2455194.01716
2455194.02147
2455194.02618
2455194.03098
2455194.03561
2455194.04057
2455194.04546
87.29
70.55
−77.76
−77.84
82.29
102.65
112.70
−57.09
86.55
107.21
117.62
−58.37
−57.70
120.17
−13.75
92.67
−114.15
−225.51
−448.40
−488.00
−289.10
−186.54
−5.48
−119.86
41.30
−134.63
19.77
25.83
50.77
143.40
−30.20
48.56
168.31
181.07
72.55
172.95
102.62
−25.28
140.26
67.47
−33.41
105.31
104.10
97.82
87.20
84.10
81.03
81.95
85.47
73.88
84.75
79.58
80.27
85.58
80.71
79.41
81.14
73.98
79.68
72.95
73.63
71.22
72.92
64.82
64.92
66.44
67.07
2.00
1.44
1.49
1.72
1.41
2.05
2.00
1.97
2.12
1.80
1.90
1.66
1.40
1.71
1.88
1.64
2.39
1.54
1.49
2.88
1.44
1.63
1.90
1.62
1.50
1.51
1.50
1.75
1.68
2.24
2.91
1.55
1.97
2.06
1.78
1.73
1.53
1.29
1.30
1.49
1.50
1.49
1.48
1.49
1.50
1.68
1.59
1.49
1.61
1.69
1.63
1.59
1.64
1.44
1.51
1.42
1.59
1.60
1.49
1.60
1.54
1.49
1.45
1.55
1.54
1.55
1.62
-- 18 --
Table 1 -- Continued
BJD
RV [m s−1]
Error [m s−1]
2455194.05048
2455194.05515
2455194.05976
2455194.06437
2455194.06915
2455194.07416
2455194.07901
2455194.08376
2455194.08858
2455194.09350
2455194.09842
2455194.10345
2455194.10848
2455194.11341
2455194.11813
2455194.12270
2455194.12732
2455194.13216
2455194.13716
2455194.17667
2455196.94719
2455197.94842
2455229.08581
2455229.87780
2455232.01621
2455251.92524
2455255.82341
2455256.97046
2455260.85979
2455284.82534
2455285.89491
2455289.81794
2455311.74995
2455312.83027
2455313.74879
2455314.80031
2455320.86712
2455321.81620
64.00
54.30
56.32
61.63
51.66
49.80
50.00
53.07
46.90
51.24
51.69
51.65
54.47
48.28
43.50
49.47
47.17
47.26
41.20
33.41
59.88
−36.39
20.75
−60.13
22.91
96.71
−93.36
54.58
39.52
−171.99
−97.21
−32.19
−418.30
−254.42
−405.93
−436.91
−514.39
−436.55
1.51
1.56
1.60
1.46
1.52
1.63
1.50
1.52
1.47
1.63
1.50
1.65
1.68
1.54
1.59
1.51
1.63
1.57
1.51
1.44
1.29
2.02
1.66
1.73
1.52
2.09
1.40
1.47
1.54
1.73
1.75
1.31
1.56
1.53
1.39
1.87
1.58
1.73
Note. -- The RV was measured relative to an ar-
bitrary template spectrum; only the differences are
significant. The uncertainty given in Column 3 is the
internal error only and does not account for "stellar
jitter."
-- 19 --
Table 2. Model Parameters for HAT-P-13
Parameter
Value
Star
Mass, M⋆ [M⊙]
Radius, R⋆ [R⊙]
Projected stellar rotation rate, v sin i⋆ [km s−1]
Planet b
Mass, Mb [MJup]
Radius, Rb [RJup]
Orbital period, Pb [d]
Planet-to-star radius ratio, Rp/R⋆
Star-to-orbit radius ratio, R⋆/a
Orbital inclination, i [deg]
Impact parameter, btra
Time of midtransit, Ttra,b [BJD]
Orbital eccentricity, eb
Argument of pericenter, ωb [deg]
eb cos ωb
eb sin ωb
Velocity semiamplitude, Kb [m s−1]
Projected spin-orbit angle, λ [deg]
Companion c
Minimum mass, Mc sin ic [MJup]
Orbital period, Pc [d]
Time of inferior conjunction, Tcon,c [BJD]
Orbital eccentricity, ec
Argument of pericenter, ωc [deg]
ec cos ωc
ec sin ωc
Velocity semiamplitude, Kc [m s−1]
Other system parameters
Angle between apsides, ωb − ωc [deg]
Velocity offset, γ [m s−1]
γ [m s−1 yr−1]
1.22+0.05
−0.10
1.559 ± 0.080
1.66 ± 0.37
0.851 ± 0.038
1.272 ± 0.065
2.916250 ± 0.000015
0.08389 ± 0.00081
0.1697 ± 0.0072
83.40 ± 0.68
0.679 ± 0.043
2, 454, 779.92976 ± 0.00075
0.0133 ± 0.0041
210+27
−36
−0.0099 ± 0.0036
−0.0060 ± 0.0069
106.04 ± 0.73
1.9 ± 8.6
14.28 ± 0.28
446.27 ± 0.22
2, 455, 312.80 ± 0.74
0.6616 ± 0.0054
175.29 ± 0.35
−0.6594 ± 0.0056
0.0543 ± 0.0038
440 ± 11
36+27
−36
−100.3 ± 2.0
17.51 ± 0.90
-- 20 --
Table 3. Predicted times of inferior conjunction for HAT-P-13c
Year
Month
Date
Hour [UT]
Julian Date
Uncertainty [d]
April
July
2010
2011
October
2012
2013 December
2015
March
26
16
4
25
16
7.3
13.9
20.4
3.0
9.6
2455312.80
2455759.08
2456205.35
2456651.62
2457097.90
0.74
0.85
1.00
1.17
1.36
|
1803.00054 | 1 | 1803 | 2018-02-20T22:07:18 | Resolved Millimeter Observations of the HR 8799 Debris Disk | [
"astro-ph.EP",
"astro-ph.SR"
] | We present 1.3 millimeter observations of the debris disk surrounding the HR 8799 multi-planet system from the Submillimeter Array to complement archival ALMA observations that spatially filtered away the bulk of the emission. The image morphology at $3.8$ arcsecond (150 AU) resolution indicates an optically thin circumstellar belt, which we associate with a population of dust-producing planetesimals within the debris disk. The interferometric visibilities are fit well by an axisymmetric radial power-law model characterized by a broad width, $\Delta R/R\gtrsim 1$. The belt inclination and orientation parameters are consistent with the planet orbital parameters within the mutual uncertainties. The models constrain the radial location of the inner edge of the belt to $R_\text{in}= 104_{-12}^{+8}$ AU. In a simple scenario where the chaotic zone of the outermost planet b truncates the planetesimal distribution, this inner edge location translates into a constraint on the planet~b mass of $M_\text{pl} = 5.8_{-3.1}^{+7.9}$ M$_{\rm Jup}$. This mass estimate is consistent with infrared observations of the planet luminosity and standard hot-start evolutionary models, with the uncertainties allowing for a range of initial conditions. We also present new 9 millimeter observations of the debris disk from the Very Large Array and determine a millimeter spectral index of $2.41\pm0.17$. This value is typical of debris disks and indicates a power-law index of the grain size distribution $q=3.27\pm0.10$, close to predictions for a classical collisional cascade. | astro-ph.EP | astro-ph | Draft version March 2, 2018
Typeset using LATEX preprint style in AASTeX61
8
1
0
2
b
e
F
0
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
5
0
0
0
.
3
0
8
1
:
v
i
X
r
a
RESOLVED MILLIMETER OBSERVATIONS OF THE HR 8799 DEBRIS DISK
David J. Wilner,1 Meredith A. MacGregor,1, 2, ∗ Sean M. Andrews,1 A. Meredith Hughes,3
Brenda Matthews,4 and Kate Su5
1Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
2Department of Terrestrial Magnetism, Carnegie Institution for Science, 5241 Broad Branch Road, Washington, DC
20015, USA
3Department of Astronomy, Van Vleck Observatory, Wesleyan University, 96 Foss Hill Drive, Middletown, CT 06459,
USA
4National Research Council of Canada, Herzberg Astronomy and Astrophysics Programs, 5071 West Saanich Road,
Victoria, BC, V9E 2E7, Canada
5Steward Observatory, University of Arizona, Tucson, AZ 85721, USA
ABSTRACT
We present 1.3 millimeter observations of the debris disk surrounding the HR 8799 multi-planet
system from the Submillimeter Array to complement archival ALMA observations that spatially fil-
tered away the bulk of the emission. The image morphology at 3.′′8 (150 AU) resolution indicates an
optically thin circumstellar belt, which we associate with a population of dust-producing planetes-
imals within the debris disk. The interferometric visibilities are fit well by an axisymmetric radial
power-law model characterized by a broad width, ∆R/R & 1. The belt inclination and orientation
parameters are consistent with the planet orbital parameters within the mutual uncertainties. The
models constrain the radial location of the inner edge of the belt to Rin = 104+8
−12 AU. In a simple
scenario where the chaotic zone of the outermost planet b truncates the planetesimal distribution,
this inner edge location translates into a constraint on the planet b mass of Mpl = 5.8+7.9
−3.1 MJup.
This mass estimate is consistent with infrared observations of the planet luminosity and standard
hot-start evolutionary models, with the uncertainties allowing for a range of initial conditions. We
also present new 9 millimeter observations of the debris disk from the Very Large Array and deter-
mine a millimeter spectral index of 2.41 ± 0.17. This value is typical of debris disks and indicates
a power-law index of the grain size distribution q = 3.27 ± 0.10, close to predictions for a classical
collisional cascade.
Keywords: circumstellar matter - stars: individual (HR 8799) - submillimeter: plan-
etary systems
[email protected]
∗ NSF Postdoctoral Fellow
2
1. INTRODUCTION
The young (30 Myr, Malo et al. 2013) and nearby (39.4 pc, van Leeuwen 2007) A-type star HR 8799
is the host of the first (and so far only) directly imaged multiple planet system. Near-infrared images
show four companions with projected separations of 14, 24, 38, and 68 AU (Marois et al. 2008,
2010) whose orbital motions have been tracked over a decade (e.g. see the compilation by Bowler
2016). Comparison of infrared photometry with standard ("hot-start") evolutionary models suggest
these companions have masses in the range of 5 − 10 MJup (Marley et al. 2012). This inference is
consistent with calculations that imply planet masses < 10 MJup for dynamical stability at the age
of the system, which also could be bolstered by a 1:2:4:8 mean motion resonance configuration (e.g.
Go´zdziewski & Migaszewski 2014; Pueyo et al. 2015; Maire et al. 2015). Analysis of self-consistent
and homogeneous astrometric measurements indicate that the planet orbits have low eccentricity
and are consistent with coplanarity (Konopacky et al. 2016). These relatively bright and wide-
separation super-Jovian planets are also a favorite target for spectroscopic observations aimed at
planetary atmosphere characterization (Barman et al. 2011; Konopacky et al. 2013; Ingraham et al.
2014; Barman et al. 2015; Bonnefoy et al. 2016; Zurlo et al. 2016).
The HR 8799 system hosts, in addition to planetary mass companions, a dusty debris disk that
was first detected by IRAS (Sadakane & Nishida 1986). The debris disk has been investigated in
detail at wavelengths from 24 − 850 µm using observations from Spitzer (Su et al. 2009), Herschel
(Matthews et al. 2014), and the JCMT (Williams & Andrews 2006; Holland et al. 2017). These
multi-wavelength observations show that the debris consists of a warm (T∼ 150 K) inner belt and a
cold (T∼ 35 K) outer belt that bracket the orbits of the directly imaged planets, plus an extended
halo of small grains that is detected out to radii beyond 1000 AU. The 850 µm photometry marks
the HR 8799 debris disk as one of the most massive known, at an estimated mass of ∼ 0.1 MEarth.
Modeling of resolved far-infrared images suggests the presence of a population of colliding plan-
etesimals underlying the cold belt that extends from about 100 to 300 AU, albeit with significant
uncertainty on the locations of these boundaries on account of insufficient angular resolution for the
inner edge and confusion with the extended halo. The relatively large radial extent of this belt cou-
pled with the young age of the system supports "planet-stirring" models to produce the collisional
debris (Mo´or et al. 2015).
Millimeter emission selectively reveals the large dust grains less affected by radiative forces that
therefore trace best the distribution of dust-producing planetesimals within debris disks (Wyatt 2006;
Wilner et al. 2011). Single dish observations of HR 8799 from the CSO at 350 µm (9′′ beam) resolved
emission from the cold belt around the star, with a tentative offset asymmetry (Patience et al. 2011).
This asymmetry was not confirmed by observations with higher signal-to-noise ratio from the JCMT
at 450 µm and 850 µm (8′′ and 13′′ beam, respectively) (Holland et al. 2017). Millimeter imaging
of the disk at much higher angular resolution has proven challenging on account of its low surface
brightness. Hughes et al. (2011) made the first millimeter interferometric detection of the disk,
using the Submillimeter Array (SMA) at 870 µm, consistent with the presence of a broad belt
(∆R/R > 1) of inner radius ∼ 150 AU. Recently, Booth et al. (2016) used the Atacama Large
Millimeter/submillimeter Array (ALMA) to image the HR 8799 system at 1.3 mm at much higher
sensitivity, although these observations recovered only a small fraction of the total disk emission.
Analysis of these ALMA data by an unconventional fitting of belt models to dirty images placed
apparently strong constraints on the structure. In particular, their analysis determined an inner edge
3
too far out to be truncated by the outermost planet b and raised the possibility of the presence of
an additional (and unseen) outer planet.
To better determine the properties of the cold belt of planetesimals in the HR 8799 system, we used
the SMA to obtain new observations at 1.3 millimeters wavelength that are sensitive to larger angular
scales than the previous millimeter interferometer studies by Hughes et al. (2011) and Booth et al.
(2016). We also obtained new observations at 9 millimeters wavelength with the Karl G. Jansky Very
Large Array (VLA), to constrain the spectral index of the disk emission. In §2, we describe the details
of these 1.3 millimeter and 9 millimeter observations. In §3, we present the 1.3 millimeter results in
the form of images and deprojected visibility functions, and provide a quantitative analysis of these
data using parametric model fits in an MCMC framework. In §4 we discuss implications of the new
millimeter results for the system geometry, the mass of planet b, and the grain size distribution. In
§5, we summarize the main conclusions.
2. OBSERVATIONS
2.1. Submillimeter Array
We observed the HR 8799 system with the SMA on Mauna Kea, Hawaii, at a wavelength near
1.3 millimeters in the subcompact configuration that provides projected baselines as short as 6 meters.
Table 1 provides a log of these observations, including the observing dates, weather conditions,
number of operational antennas, and the useable hour angle range.
In this close-packed antenna
configuration, the lower elevation limit for observations is 33◦, to prevent antenna collisions. The
phase center was set at α = 23h07m28.72, δ = +21◦08′03.′′3 (J2000), i.e. the J2000 star position
uncorrected for interim proper motion. Two units of the SWARM digital correlator, still under
construction at the time, were available; together these spanned bands ±(4 − 12) GHz from the
LO frequency of 225.5 GHz, with uniform channel spacing 140 kHz (∼0.18 km s−1). This setup
provided a total of 16 GHz of continuum bandwidth, as well as simultaneous coverage of the 12CO
J=2-1 (230.53800 GHz) and 13CO J=2-1 (220.39868 GHz) spectral lines in the upper and lower
sidebands, respectively. The primary beam FWHM size of the 6-meter diameter array antennas of
55′′(ν/230 GHz) set the useable field of view.
The basic observing sequence consisted of a loop of 2 minutes each on the quasars 3C454.3 (∼ 13 Jy,
6◦ away) and J2232+117 (∼ 5 Jy, 12◦ away) and 10 minutes on the target source HR 8799. Passband
calibration was obtained with observations of the strong sources 3C454.3 and Uranus. The absolute
flux scale was set using observations of Uranus obtained in each track, with 10% overall accuracy.
All of the basic calibration was performed using standard procedures in the MIR software package.
Continuum visibilities were output in 30 second scans spanning 4 GHz widths, centered at 215.5,
219.5, 231.5, and 235.5 GHz. Fourier inversion for continuum and spectral line imaging and clean
deconvolution were done in the CASA software package (version 4.7.2).
2.2. Atacama Large Millimeter/submillimeter Array
We retrieved observations of HR 8799 from the ALMA archive that were made in Band 6 in 2015
January. Details of these observations are described in Booth et al. (2016). We calibrated the five
individual scheduling block executions using the associated CASA 4.3.1 reduction scripts. In brief, the
correlator setup consisted of four spectral windows with width 2 GHz and channel spacing 16 MHz,
centered at 216, 218, 231 and 233 GHz, respectively. The observations included baseline lengths
from 15 to 349 meters. The primary beam FWHM size of the 12-meter diameter array antennas
4
Table 1. SMA 1.3 Millimeter Observations of HR 8799
Observation
# of
Projected
τatm
H.A.
Date
Antennas Baselines (m) 225 GHz
range
2016 Aug 17
2016 Aug 27
2016 Aug 29
2016 Aug 30
2016 Sep 06
2016 Sep 21
7
8
8
8
8
7
7 − 69
6 − 69
6 − 69
6 − 69
6 − 69
6 − 45
0.08
0.08
0.10
0.07
0.09
0.08
−3.9, +3.7
−4.0, +3.8
−1.7, +3.7
−0.8, +3.7
−3.8, +3.7
−3.6, +3.6
of 27′′(ν/230 GHz) set the useable field of view. We used the CASA task statwt to homogenize the
visibility weights among scheduling blocks and note that the execution on 2015 January 03 obtained
in the best weather conditions dominates the sensitivity budget. Continuum visibilities were output
for analysis after flagging the channels with CO line emission, averaging in time to 30 seconds, and
averaging in frequency to 2 GHz bandwidth. Imaging was carried out with the clean task in CASA
(version 4.7.2).
2.3. Very Large Array
We observed the HR 8799 system with the VLA at 9 mm wavelength in the most compact D
configuration. Observations with the 27 array antennas were executed in 2 hour scheduling blocks
on 4 dates in early 2017: Feb 16, Mar 04, 11, and 13. The baselines ranged from 0.04 to 1.03
km. The weather conditions were very good, with phase noise measurements from the Atmospheric
Phase Interferometer (on a 300 m baseline at 11.7 GHz) ranging from 3.3 to 7.4◦. The phase center
was set to the same position as for the SMA observations. The correlator was used to obtain the
maximum continuum bandwidth of 8 GHz, comprised of 4 bands centered at 30, 32, 34, and 36
GHz, for two polarizations. The primary beam FWHM size of the 25-meter diameter array antennas
is approximately 85′′(ν/30 GHz). The observing sequence interleaved 1 minute observations of the
complex gain calibrator 3C454.3 with 5 minute observations on HR 8799. Passband calibration
was obtained with observations of 3C454.3. The absolute flux scale was set with observations of the
standard calibrator 3C48 in each execution, with an estimated accuracy of 10%. The basic calibration
was done using VLA pipeline processing in the CASA package (version 4.7.2) followed by imaging with
clean task.
3. RESULTS AND ANALYSIS
3.1. Continuum Emission
Figure 1 shows 1.3 mm continuum images of HR 8799 from the SMA (left), ALMA (middle), and
the combination of SMA and ALMA (right). These images were obtained with natural weighting, and
those including SMA data use a 4′′ FWHM Gaussian taper to improve surface brightness sensitivity
at the expense of angular resolution. The ALMA image is comparable to Figure 1 of Booth et al.
5
Figure 1.
Images of 1.3 mm continuum emission from the HR 8799 debris disk from the SMA (left), ALMA
(center), and the SMA and ALMA combined (right). All of these images were made with natural weighting, and those
including SMA data make use of a 4′′ FWHM Gaussian taper. Contour levels are in steps of 2× the measured rms
values are 180 µJy/beam, 16 µJy/beam, and 30 µJy/beam for the three images, respectively. The beam sizes are
shown by the ellipses in the lower left corner of each image and are 6.′′1 × 5.′′6, 1.′′7 × 1.′′2, and 3.′′8 × 3.′′7, respectively.
The blue star symbol indicates the stellar position, and the dashed gray line shows the best-fit belt position angle (see
§3.3).
(2016), with a beam size of ∼ 1.′′5, and it shows a noisy ring of emission surrounding the star. For the
SMA image, the beam size is ∼ 4× larger in each dimension than for the ALMA image, and the noise
level is an order of magnitude higher, but the sensitivity to larger angular scales provided by the
shorter baselines enables a much improved representation of the radially extended belt of emission
around the star. The combination of SMA and ALMA data results in an even better image of this
circumstellar structure. In the SMA and ALMA combined image shown with a ∼ 3.′′8 beam, the two
peaks visible on either side of the star are the signatures of limb brightening along the major axis of
an inclined and optically thin emission belt.
Figure 2 shows the deprojected visibility function for the emission from the SMA and ALMA,
obtained by averaging the real and imaginary parts of of the complex visibilities in concentric annular
bins, adopting the belt geometry derived in §3.3. This view of the visibilities provides some insight
into the image properties in Figure 1. The shorter baselines from the SMA capture the steep rise in
visibility amplitude at low spatial frequencies that are missed by ALMA, and thus sample better the
emission responsible for the extended disk morphology. The visibilities from the two telescopes are
consistent where they overlap at higher spatial frequencies that sample the small scale structure. The
imaginary parts of visibilities are all consistent with zero, as expected for an axisymmetric structure.
Figure 3 shows a VLA 9 millimeter continuum image of the HR 8799 system, obtained with natural
weighting and a 5′′ FWHM Gaussian taper, which gives a beam size of 6.′′2 × 5.′′8 and rms noise of
4.3 µJy/beam. Significant emission is detected from an extended region centered on the star, but
with a low peak signal-to-noise ratio of only 3 − 4 per beam. No clear disk or belt morphology is
6
Figure 2. The real and imaginary parts of the 1.3 mm visibilities, averaged in bins of deprojected (u, v) distance
from the star, for data from the SMA (black filled circles) and ALMA (gray open circles). The solid blue line depicts
the best-fit power law axisymmetric belt model to the combined dataset (see §3.3). The imaginaries are consistent
with zero as is expected for an axisymmetric belt.
evident in the emission at this sensitivity level. The emission peak near (+10′′, −10′′) in the image
could be the result of a noise fluctuation, or perhaps a faint and unrelated background source.
3.2. Line Emission
Emission from the 12CO and 13CO J=2-1 lines were clearly detected by the SMA observations
over the narrow vLSR range −6 to −4 km s−1, close to the stellar velocity of −4.6 km s−1 in
this reference frame. Single dish observations of the 12CO J=3-2 line with a 15′′ beam show ex-
tended emission, concentrated in a filament that extends from the north-northeast of HR 8799 to the
south-southwest, most likely affiliated with the nearby MBM 53-55 complex of high latitude clouds
(Williams & Andrews 2006; Su et al. 2009). The submillimeter dust continuum emission observed
by Herschel also shows traces of this large scale feature (Matthews et al. 2014). Like previous at-
tempts to use interferometers to image the molecular emission in this region, the extended structure
evident in lower angular resolution single dish maps (Su et al. 2009) remains poorly sampled and
highly problematic to reconstruct, even with the lower spatial frequency information newly obtained
7
Figure 3.
Image of the 9 millimeter continuum emission from the HR 8799 debris disk from the VLA obtained with
natural weighting and a 5′′ FWHM Gaussian taper. Contour levels are ±3, 5× the measured rms of 4.3 µJy/beam.
The 6.′′2 × 5.′′8 beam size is indicated by the ellipses in the lower left corner and the stellar position by the blue star
symbol.
with the SMA. Regardless of the choice of visibility weighting scheme and clean deconvolution pa-
rameters, all of our attempts at imaging the line emission show pronounced artifacts. Emission from
the 13CO J=2-1 line, with substantially lower optical depth than the main isotopolgue, offers the
best chance to reveal structure associated with the HR 8799 system. Figure 4 shows a set of 13CO
channel maps obtained with natural weighting and a 4′′ FWHM Gaussian taper, resulting in a ∼ 6.′′5
beam. Among the positive and negative corrugations indicative of missing flux, the most significant
emission is located away from the star, and there is no clear evidence for any systematic motions
that might be associated with rotation of gas in the debris disk. The linewidth is consistent with the
typical turbulent broadening found in small molecular clouds. The apparent association in space and
velocity between HR 8799 and this extended molecular emission is intriguing but remains mysterious.
SMA channel maps of 13CO J = 2 − 1 emission for HR 8799. For all panel, the contours are in steps
Figure 4.
of ±2× the rms noise level of 100 mJy/beam. The ellipse in the lower left corner of the leftmost panel indicates the
7.′′0 × 6.′′1 synthesized beam size resulting from natural weighting and a 4′′ Gaussian taper. The LSR velocities are
labeled in the upper left corner of each panel.
3.3. SMA/ALMA Disk Model Fits
To provide a quantitative characterization of the 1.3 millimeter emission from the HR 8799 disk,
we make the simple assumption that the structure can be represented by an axisymmetric and ge-
ometrically thin belt, and we model the observations using the visibility fitting procedure described
in MacGregor et al. (2013). The surface brightness of the belt is assumed to be a radial power law,
8
Iν ∝ rx, between an inner and outer radius, Rin and Rout, respectively. The surface brightness
power law index, x, encapsulates both the radial surface density and temperature profiles of the
emitting grains. The total flux density of the belt is normalized to Fbelt = R IνdΩ. We also fit
for the belt geometry, specifically an inclination, i (where 0◦ is face-on), and a position angle, P A
(measured East from North). For comparison with observed visibilities, the model emission is multi-
plied by the appropriate primary beam response of each telescope. Since imaging with both Herschel
(Matthews et al. 2014) and JCMT (Holland et al. 2017) shows the presence of a compact background
source to the northwest of the disk, we also include an unresolved point source at this location in
the model, described by a total flux, Fpt, and an offset from the phase center of the observations,
{∆αpt, ∆δpt}. When modeling the ALMA data alone, this additional background source is omitted
on account of primary beam attenuation at the offset location.
This simple two-dimensional parametric model for the disk emission is incorporated into a Markov
Chain Monte Carlo (MCMC) framework. For each model image, we calculate synthetic visibilities
using vis sample1, a python package for visibility sampling that improves on the sampling and
interpolation of the uvmodel task in the Miriad software package, and determine a χ2 likelihood
function, lnL = −χ2/2, that incorporates the statistical weights of the visibilities. We make use of the
emcee package (Foreman-Mackey et al. 2013) to sample the posterior probability distribution of the
data conditioned on the model, and determine the best-fit parameter values and their uncertainties.
We explore the posterior with 50,000 iterations (100 walkers and 5, 000 steps each) and check that
the Gelman-Rubin statistic (Gelman et al. 2014) is R < 1.1 for all parameters to ensure convergence.
We first fit the SMA observations and the ALMA observations separately, to examine the constraints
provided by each dataset alone, and then fit the observations from both telescopes simultaneously, to
examine the joint constraints. Table 2 lists the best-fit parameters from each of these fits, together
with their 68% uncertainties determined from the marginalized posterior probability distributions.
For modeling the SMA observations alone, we adopt uniform priors for all parameters, in particular
0 < Rin < Rout, Fbelt > 0, and −3 < x < 3. When modeling the ALMA observations alone,
informative constraints were not obtained for the outer radius or total flux parameters using these
assumptions.
In effect, the ALMA data allow the outer radius parameter to extend freely, with
ever increasing total flux. This most likely results because the ALMA primary beam is comparable
in size to the HR 8799 emission region, and the ALMA observations provide very little data at
spatial frequencies inside the first null of the visibility function to limit the total disk emission. To
obtain constraints on these parameters from the ALMA data alone, we must impose strong priors
on both Rout and Ftot. As an example, we implemented a Gaussian prior on the total belt flux
density, 3.0 ≤ Ftot ≤ 4.0 mJy, motivated by submillimeter observations (Matthews et al. 2014)
that predict a flux density of ∼ 3.5 mJy at 1.3 mm assuming a millimeter spectral index of 2.5,
typical of debris disks (MacGregor et al. 2016). We note that the fit to the SMA data alone yields
a higher flux density of 5.13+0.34
−1.01 mJy, but still comparable within the uncertainties. In addition, we
implemented a logistic function f (x) = 1/(1 + e−k(x−x0)) prior for the outer radius with k = −10 and
x0 = 512 AU, the primary beam diameter. With these assumptions, fitting the ALMA observations
alone yields Ftot = 3.5 ± 0.5 mJy and Rout = 497.(+2., −150) AU, effectively recovering the priors.
We suspect that the ∼ 10% uncertainties on the disk parameters quoted by Booth et al. (2016)
1 vis sample is publicly available at https://github.com/AstroChem/vis_sample or in the Anaconda Cloud at
https://anaconda.org/rloomis/vis_sample
9
from an image-based analysis of the ALMA observations are underestimated, perhaps in part due
to unaccounted covariance associated with large scale emission. By contrast, modeling the SMA
and ALMA observations simultaneously does not require strong priors on the disk parameters to
obtain meaningful constraints. Figure 5 shows the posterior probability distributions for the model
parameters that result from fitting the SMA and ALMA observations together. For these combined
fits, the small proper motion of HR 8799 between SMA and ALMA observing epochs (0.′′20) was
ignored, given its insignificance at the low angular resolution of the SMA observations.
Figure 6 shows the combined SMA and ALMA image from Figure 1 together with images of the
best-fit model at full resolution, the best-fit model imaged in the same way as the data, and an
image made from the observed visibilities after subtracting the best-fit model visibilities. This broad
(∆R/R & 1) inclined belt model reproduces very well the main features of the observations, including
the inner depression and limb-brightened regions in the northeast and southwest. The goodness-of-fit
is demonstrated by the noise-like residual image. Notably, the imaged residuals reveal no significant
azimuthal brightness asymmetries, as might arise from the presence of resonant clumps, or an offset
of the star from the center of the disk. If any such asymmetries are present in the millimeter emission
from the HR 8799 disk, observations with a significantly higher signal-to-noise ratio will be required
to reveal them. This outer belt of large fractional width joins a growing list other young debris
disks with similar characteristics, e.g. HD 95086 (Su et al. 2017), that are much broader than the
classical Kuiper Belt of our Solar System (especially the low excitation "kernel" component discussed
by Bannister et al. 2016).
3.4. VLA 9 millimeter Flux Estimate
In the VLA image, the disk is detected only at the ∼ 3σ per beam level, and we do not apply
the same MCMC modeling technique to these data. For simplicity, we measure the flux density
in CASA within a circular aperture centered at the location of the disk, assuming the diameter of
the disk determined from the combined modeling of the SMA and ALMA data. We determine the
uncertainty by measuring the flux density in the same circular aperture at ten random locations
in the image and taking the average. The measured total flux density of the disk from our VLA
observations is F9mm = 32.6 ± 9.9 µJy (correction for primary beam attenuation is negligible). We
adopt this image-based approach instead of fitting to the millimeter visibilities due to the difficulty
in selecting a function that adequately reproduces the complex structure of the observed surface
brightness profile.
4. DISCUSSION
By modeling new and archival resolved millimeter observations, we have made an improved charac-
terization of the planetesimal belt within the debris disk surrounding the HR 8799 planetary system.
This broad emission belt spans from ∼ 100−360 AU, roughly consistent with previous modeling of the
spectral energy distribution (Su et al. 2009) and fitting far-infrared emission images (Matthews et al.
2014). The 1.3 millimeter interferometric dataset from SMA and ALMA together recovers the bulk
of the disk flux and provides high angular resolution, and we used this combined dataset to fit for
basic structural parameters. We next consider implications. In particular, we examine the belt ge-
ometry constraints in the context of coplanarity and alignment of the planetary system (§4.1), the
relationship between the belt inner edge and the outermost planet b (§4.2), ramifications of the radial
10
Figure 5. Corner plot summary of parameter values from the MCMC belt modeling of the combined SMA and
ALMA data. The panels on the diagonal show the 1-D histogram for each model parameter obtained by marginalizing
over the other parameters, with a dashed vertical line to indicate the best-fit value. The off-diagonal panels show 2-D
projections of the posterior probability distributions for each pair of parameters, with contours to indicate 1σ (red)
and 2σ (grey) regions.
surface density profile for collisional excitation mechanisms (§4.3), and the millimeter spectral index
connections to debris disk collisional models (§4.4).
11
Figure 6.
(left) The combined SMA and ALMA image of the HR 8799 debris disk. (center left) The best-fit model
to the combined datasets displayed at full resolution (pixel size 0.′′5 = 2 AU at 39.4 pc). (center right) The best-fit
combined model convolved with the synthesized beam. (right) The residuals obtained from imaging the best-fit model
subtracted from the data. In all panels except the full resolution model, contours are in steps of 2 × 30 µJy/beam, the
rms noise level. The blue star marks the stellar position, the dashed gray line indicates the best-fit position angle of
the disk, and the dashed gray ellipse in the leftmost panel shows the 3.′′8 × 3.′′7 synthesized beam.
Table 2. Best-fit Model Parameters
Parameter
Description
SMA
ALMAa
SMA+ALMA
Fbelt
Rin
Rout
x
i
P A
Fpt
∆αpt
Total belt flux density [mJy]
5.13(+0.34, −1.01)
3.50 ± 0.50
3.44(+0.31, −0.56)
Belt inner edge [AU]
Belt outer edge [AU]
106.(+19., −28.)
115.(+16., −17.)
104.(+8., −12.)
370.(+42., −40.)
497.(+22., −150.)
361.(+16., −18.)
Surface brightness power law index
−0.56(+0.31, −0.83) −0.64(+0.40, −0.46) −0.38(+0.47, −0.47)
Disk inclination [◦]
Disk position angle [◦]
35.6(+5.6, −4.4)
34.3(+5.9, −6.9)
32.8(+5.6, −9.6)
41.6(+9.6, −7.8)
43.7(+12.2, −12.4)
35.6(+9.4, −10.1)
background source flux density [mJy]
0.47(+0.11, −0.17)
background source RA offset [′′]
−12.9(+0.2, −0.9)
· · ·
· · ·
∆δpt
aAdopting strong priors on the total belt flux density and belt outer edge parameters (see §3.3).
background source DEC offset [′′]
17.8(+0.4, −0.6)
· · ·
0.45(+0.12, −0.15)
−12.2(+0.8, −0.8)
17.5(+0.9, −0.5)
4.1. System Geometry
The model fits to the millimeter emission provide new estimates of the inclination and orienta-
tion of the planetesimal belt that can be compared directly to constraints on the orbital elements
of the planets. Formation of planets in a thin disk will generally lead to orbital coplanarity and
alignment, unless upset by dynamical events. Since the orbital periods of the HR 8799 outer planets
span centuries, only small arcs of their motions around the star have been measured in the years
since discovery, leaving significant uncertainties in their orbital elements. Early stability calculations
indicated departures from a face-on geometry (Fabrycky & Murray-Clay 2010), and most analy-
ses have favored orbits with inclinations of 10 to 30◦ (e.g. Currie et al. 2012; Esposito et al. 2013;
12
Pueyo et al. 2015; Wertz et al. 2017). A recent analysis of carefully calibrated astrometric monitor-
ing data that minimizes systematic errors generally indicate more highly inclined orbits, specifically
38◦ ±7◦ for the outermost planet b, and no evidence for any measureable departures from coplanarity
(Konopacky et al. 2016). Astroseismology suggests a higher inclination angle for the star, & 40◦, as
well (Wright et al. 2011). These results are all compatible with the disk inclination of 32.◦8+5.6
−9.6 derived
from the combined SMA and ALMA data.
There is some tension between the millimeter result and the lower (and more precise) disk inclination
of 26◦ ± 3◦ determined from the ellipticity of the outer edge of far-infrared emission in Herschel
images (Matthews et al. 2014). As noted by Booth et al. (2016), the discrepancy between millimeter
and far-infrared disk inclinations may have a physical origin in size-dependent dust dynamics, or in
complexities in disk structure not captured by the simple belt models. Taking the millimeter emission
as the best tracer of the dust-producing planetesimals, the various inclination determinations are
compatible with coplanarity with the planetary orbits, and with the star, although the uncertainties
remain large enough to preclude drawing a firm conclusion.
If the planetesimal belt and planet orbits are coplanar, as suggested by the derived inclination angle,
then we would also expect alignment of the position angle of the belt and the argument of ascending
node of the planet orbits, Ω. The Konopacky et al. (2016) astrometric analysis indicates Ω lies in the
range of 40 to 70◦ for low eccentricity solutions, for all four of the planets. The disk position angle
of 35.◦6+9.4
−10.1 determined from the joint analysis of the SMA and ALMA data is compatible with the
low end of this range. Like the inclination, the disk position angles derived from millimeter and far-
infrared emission are discrepant, with the latter yielding a higher value of 64◦ ± 3◦ (Matthews et al.
2014). Again, there may be a physical origin for this difference. The overlap of the millimeter
constraint on the planetesimal belt position angle and the planetary argument of ascending node
provides weak additional support for coalignment of these components.
4.2. Dynamical Constraints on the Mass of Planet b
Because the wide-separation HR 8799 planets are relatively accessible to optical and near-infrared
observations, they have become key benchmark objects for models of planetary atmospheres and
evolution. Evolutionary models are highly degenerate in planet mass, luminosity and age, however,
in part because the imprint of initial conditions can persist for 100's of Myr (Marley et al. 2007).
Depending on initial thermal content, giant planets can exhibit a wide range of luminosities.
In
principle, independent constraints on planet mass from dynamics offer the potential to break these
model degeneracies. Such constraints have the potential to help distinguish between planet formation
models, in particular hot-start models where most entropy is retained, e.g. by gas that collapses
directly to form planets, from cold-start models where most entropy is lost, e.g. by gas cooling in
core accretion. Notably, the core accretion mechanism is challenged to form 5 − 10 MJup planets in
situ at the large orbital radii of the HR 8799 planets (Kratter et al. 2010). The observed luminosities
of the HR 8799 planets cannot be reproduced by the most extreme cold-start models proposed by
Marley et al. (2007). However, they are compatible with the intermediate "warm-start" models
discussed by Spiegel & Burrows (2012) and Marleau & Cumming (2014), since luminosity increases
with both mass and initial entropy (for a given age). These models would imply much higher planet
masses than hot-start models by up to a factor of two (or perhaps more). For planet b, a fully
self-consistent solution for the planet properties remains elusive once spectroscopic constraints are
combined with infrared photometry, likely due to limitations of the models (Marley et al. 2012).
13
Dynamical constraints on planet masses would be valuable, but they are generally difficult to obtain
for wide-separation planets that are impractical to monitor over many orbital periods.
For HR 8799, analysis of the long-term stability of the system offers one form of dynamical
mass constraint on the planets. The HR 8799 multi-planet system was investigated numerically
by Gotberg et al. (2016) who found that nominal models with 5–7 MJup planets are unstable on
timescales much shorter than the system age. One natural solution to this problem is if the planet
masses are substantially lower than estimated from infrared observations. Lower mass planets have
larger separations in terms of Hill radii, and the system stability persists for much longer times.
If standard cooling models underpredict the planet luminosities, as some calculations suggest (e.g.
Mordasini 2013), then lower planet masses – more like 2–3 MJup – might be accommodated. Alter-
natively, much higher planet masses can be viable if the orbits are stabilized by resonant lock. The
configuration of the HR 8799 planet orbits appear to be consistent with a 1:2:4:8 resonance (e.g.
Reidemeister et al. 2009; Konopacky et al. 2016).
The truncation of the inner edge of the planetesimal disk by the gravity of the outermost planet b
offers another form of dynamical mass constraint. The separation of the planet from the disk is
sensitive to the planet mass, as the planet clears a chaotic zone around itself due to resonance
overlap (e.g. Quillen 2006; Chiang et al. 2009). For the simple case of circular and coplanar orbits,
Wisdom (1980) derived an analytic formula for the width of the chaotic zone exterior to the planetary
orbit, ∆a = Capl(Mpl/M∗)2/7, where C is a scaling coefficient of 1.3, apl is the semi-major axis of
the planet, and Mpl and M∗ are the mass of the planet and the star, respectively. Other work that
adopts a slightly different definition of resonance width gives a different scaling factor (e.g. 1.4,
Malhotra 1998). The width of the chaotic zone for more complex configurations has been the subject
of numerous investigations, considering additional factors like the eccentricity of the planet orbit
(Quillen & Faber 2006; Regaly et al. 2017), and eccentricities of the planetesimal orbits with the
disk (Mustill & Wyatt 2012; Pearce & Wyatt 2014). The Pearce & Wyatt (2014) study formulates
a convenient expression for the location of the disk inner edge relative to the width of the chaotic
zone, Rin = apl + 5apl(Mpl/3M∗)1/3, taken in the limit of zero eccentricity.
If we adopt the Pearce & Wyatt (2014) formula, then we can translate the radial location of the
inner edge of the millimeter belt, Rin = 104+8
−12 AU, directly into a constraint on the mass of the
outermost planet b. Figure 7 shows the result of this translation of the inner radius posterior
distribution from our modeling, assuming planet b truncates the belt. For this calculation, we adopt
the planet semi-major axis apl = 68 AU and stellar mass M∗ = 1.56 M⊙ from dynamical analysis
(Soummer et al. 2011), consistent with optical photometry and spectral synthesis (Gray & Kaye
1999). This procedure gives Mpl = 5.8+7.9
−3.1 MJup for the planet b mass, compatible with standard
planet cooling models, though the errors are still large enough to encompass both hot-start and a
range of warm-start initial conditions. This new determination improves on previous attempts that
use an inner belt radius estimated from the spectral energy distribution (Su et al. 2009), or from the
ALMA data alone (Booth et al. 2016), both of which suffer from substantial systematic uncertainty.
Given this result, there is no need to posit the presence of an additional unseen planet orbiting beyond
planet b, or variations in the orbit of planet b, to account for the disk truncation. Since this planet
mass constraint scales very steeply with the disk-planet separation, even a modest improvement in
the constraint on the radial location of the inner edge of the millimeter belt has the potential to
significantly reduce the uncertainty on the planet b mass estimate. Deeper observations of HR 8799
14
could be readily obtained with the SMA, or the Atacama Compact Array (ACA), at the appropriate
spatial frequencies, to improve the constraint. A better planet b mass estimate obtained in this way
could provide important feedback into the system dynamics, planet cooling models, and formation
scenarios.
Figure 7. Constraint on the mass of planet b obtained by assuming this planet truncates the inner edge of the
HR 8799 millimeter belt. To generate this figure, we have translated the inner radius posterior distribution from our
MCMC modeling of the millimeter emission shown in Figure 5. The solid gray line indicates the best-fit results, and
the dashed gray lines indicate the 1σ uncertainties.
4.3. Stirring Mechanisms
The planetesimals within debris disks like HR 8799 must be "stirred" in order to incite destructive
collisions (Wyatt 2008). The two leading mechanisms are self-stirring by an outwardly moving front
of planetoid formation and growth (Kenyon & Bromley 2002) and planet-stirring by the gravita-
tional influence of interior giant planets (Mustill & Wyatt 2009). Given the young age of HR 8799,
Mo´or et al. (2015) concluded that the disk is most likely planet-stirred, since the long timescales
needed to form Pluto-sized bodies at 100's of AU are not compatible with the large extent of the disk
that is inferred from far-infrared imaging. The extent of the planetesimal belt more directly traced
by millimeter emission supports this conclusion. Moreover, it seems likely that the giant planets
responsible for stirring the debris disk are detected directly.
An additional observable that may help to discriminate between stirring scenarios is the radial
distribution of planetesimals. Self-stirred models produce an extended planetesimal disk with an
outwardly increasing surface density (e.g., Σ ∝ r+7/3 in the models of Kennedy & Wyatt 2010). By
contrast, standard steady-state debris disk models tend to produce surface density gradients that
decrease with radius (e.g. Krivov et al. 2006). High quality resolved millimeter observations of a
15
handful of debris disks suggest either rising surface density profiles (MacGregor et al. 2015) or hint
at more complex radial structures (e.g. multiple cold belts in HD 107146, Ricci et al. 2015). From
our model fits for HR 8799, the best-fit power law index for the millimeter emission radial profile is
x = −0.38 ± 0.47. If we assume a radial temperature profile T (r) ∝ r−0.5 to approximate radiative
equilibrium with stellar heating, then this implies a surface density profile Σ(r) ∝ r0.12±0.47. While
the uncertainties are still significant, this best-fit shallow dependence of planetesimal surface density
on radius is not predicted in self-stirring scenarios. This result also favors planet-stirring as the basic
collisional excitation mechanism for this radially extended debris disk.
4.4. Grain Size Distribution
The spectral index of dust emission at millimeter wavelengths from debris disks encodes information
on the grain size distribution that can be used to assess collisional models (Ricci et al. 2012). For
a power law grain size distribution, n(a) ∝ a−q, the reference model is the classical steady state
collisional cascade where q = 7/2 (Dohnanyi 1969). Lower or higher values of q may be obtained by
varying the strength and velocities of colliding bodies (Pan & Sari 2005; Pan & Schlichting 2012),
and are found in time-dependent numerical models of debris disk evolution (e.g. Schuppler et al.
2015).
For HR 8799, we use the best-fit total flux density from the combined SMA and ALMA models at
1.3 millimeters and the VLA flux density measured at 9 millimeters to obtain a millimeter spectral
index of αmm = 2.41 ± 0.17. Given a grain composition model, αmm provides a constraint on the size
distribution of the emitting grains, in particular on the power-law index q = (αmm − αPl)/βs + 3. In
this expression, valid for 3 < q < 4, βs is the dust opacity power law index in the small grain limit,
1.8 ± 0.2 for interstellar grain materials (Draine 2006), and αPl is the spectral index of the Planck
function, 1.92 ± 0.10 (the Rayleigh-Jeans limit modified by a small correction factor from the 39.7 K
dust temperature assumed for HR 8799). Given the measured millimeter fluxes for HR 8799, the
resulting grain size distribution power-law index is q = 3.27 ± 0.10. This result is consistent with the
weighted mean value of q = 3.36 ± 0.02 determined by MacGregor et al. (2016) for a sample of 15
debris disks observed over a similar wavelength range. This grain size distribution is most consistent
with a classical collisional cascade, with no need to invoke additional complexities in the collisional
model.
5. CONCLUSIONS
We have made new observations of millimeter dust continuum emission from the debris disk that
surrounds the four directly imaged planets in the HR 8799 system. SMA observations at 1.3 millime-
ters complement archival ALMA observations by adding measurements at lower spatial frequencies
to better sample extended emission. By fitting simple parametric disk models to these data in an
MCMC framework, we obtain constraints on the structural parameters of the millimeter emission.
The main conclusions from this analysis are:
1. The 1.3 mm emission morphology imaged at 3.′′8 (150 AU) resolution is consistent with a
broad (∆R/R & 1), axisymmetric, inclined belt centered on the star. These data provide
no evidence for the presence of clumps or other azimuthal asymmetries that would betray
dynamical sculpting of the disk by the gravity of the interior orbiting planets.
16
−9.6, and position angle, 35.◦6+9.4
2. Model fits to the combined SMA and ALMA visibilities constrain the millimeter belt inclina-
tion, 32.◦8+5.6
−10.1. These values overlap with estimates of the orbital
inclination and angle of ascending node of the planet orbits determined by astrometric obser-
vations, and they are consistent with a coplanar configuration of the disk and planetary orbits
within the (still large) mutual uncertainties.
3. The best-fit inner edge of the millimeter emission belt, Rin = 104+8
−12 AU, provides an inde-
pendent dynamical constraint on the mass of the outermost planet b of 5.8+7.9
−3.1 MJup under the
assumption that the chaotic zone of this planet truncates the planetesimal disk. This dynamical
mass estimate is commensurate with those obtained from the planet luminosity and standard
hot-start evolutionary models, although the uncertainties allow for a range of initial thermal
content.
4. Flux density measurements at 1.3 and 9 millimeters from the SMA and VLA, respectively,
give a spectral index of 2.41 ± 0.17, which implies a grain size distribution power-law index
of q = 3.27 ± 0.10. This value is consistent with the weighted mean value of q determined
from millimeter observations of other debris disks and close to predictions for a steady state
collisional cascade.
The millimeter observations presented here have improved our view of the large scale spatial dis-
tribution of the dust-producing bodies in the debris disk surrounding the HR 8799 planetary system.
However, these observations still lack the sensitivity needed to reveal any details of planet-disk inter-
actions, or to obtain dynamical constraints on planet mass that usefully discriminate between planet
evolutionary models. Deeper millimeter observations should be pursued to obtain better constraints
on the planetesimal belt morphology, and to determine if more a sophisticated approach is needed
for modeling and interpretation.
M.A.M. acknowledges support from the National Science Foundation under Award No. 1701406.
The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory
and the Academia Sinica Institute of Astronomy and Astrophysics and is funded by the Smith-
sonian Institution and the Academia Sinica. This paper makes use of the following ALMA data:
ADS/JAO.ALMA#2012.1.00482.S. ALMA is a partnership of ESO (representing its member states),
NSF (USA) and NINS (Japan), together with NRC (Canada), MOST and ASIAA (Taiwan), and
KASI (Republic of Korea), in cooperation with the Republic of Chile. The Joint ALMA Observa-
tory is operated by ESO, AUI/NRAO and NAOJ. The National Radio Astronomy Observatory is
a facility of the National Science Foundation operated under cooperative agreement by Associated
Universities, Inc.
REFERENCES
Bannister, M. T., Kavelaars, J. J., Petit, J.-M., et
Barman, T. S., Konopacky, Q. M., Macintosh, B.,
al. 2016, AJ, 152, 70
& Marois, C. 2015, ApJ, 804, 61
Barman, T. S., Macintosh, B., Konopacky, Q. M.,
Bonnefoy, M., Zurlo, A., Baudino, J. L., et al.
& Marois, C. 2011, ApJ, 733, 65
2016, A&A, 587, A58
Booth, M., Jord´an, A., Casassus, S., et al. 2016,
Maire, A.-L., Skemer, A. J., Hinz, P. M., et al.
17
MNRAS, 460, L10
Bowler, B. P. 2016, PASP, 128, 102001
Chiang, E., Kite, E., Kalas, P., Graham, J. R., &
Clampin, M. 2009, ApJ, 693, 734
Currie, T., Fukagawa, M., Thalmann, C.,
Matsumura, S., & Plavchan, P. 2012, ApJ, 755,
L34
Dohnanyi, J. S. 1969, JGR, 74, 2531
Draine, B. T. 2006, ApJ, 636, 1114
Esposito, S., Mesa, D., Skemer, A., et al. 2013,
A&A, 549, A52
Fabrycky, D. C., & Murray-Clay, R. A. 2010, ApJ,
710, 1408
Foreman-Mackey, D., Hogg, D. W., Lang, D., &
Goodman, J. 2013, PASP, 125, 306
Gelman, A., Carlin, J. B., Stern, H. S., et al. 2014,
Bayesian data analysis, Vol. 2 (CRC press Boca
Raton, FL)
Go´zdziewski, K., & Migaszewski, C. 2014,
MNRAS, 440, 3140
Gotberg, Y., Davies, M. B., Mustill, A. J.,
Johansen, A., & Church, R. P. 2016, A&A, 592,
A147
Gray, R. O., & Kaye, A. B. 1999, AJ, 118, 2993
Holland, W. S., Matthews, B. C., Kennedy, G. M.,
et al. 2017, MNRAS, 470, 3606
Hughes, A. M., Wilner, D. J., Andrews, S. M., et
al. 2011, ApJ, 740, 38
Ingraham, P., Marley, M. S., Saumon, D., et al.
2014, ApJ, 794, L15
Kennedy, G. M., & Wyatt, M. C. 2010, MNRAS,
405, 1253
Kenyon, S. J., & Bromley, B. C. 2002, ApJ, 577,
L35
Konopacky, Q. M., Barman, T. S., Macintosh,
B. A., & Marois, C. 2013, Science, 339, 1398
Konopacky, Q. M., Marois, C., Macintosh, B. A.,
et al. 2016, AJ, 152, 28
Kratter, K. M., Murray-Clay, R. A., & Youdin,
A. N. 2010, ApJ, 710, 1375
Krivov, A. V., Lohne, T., & Sremcevi´c, M. 2006,
A&A, 455, 509
MacGregor, M. A., Wilner, D. J., Rosenfeld,
K. A., et al. 2013, ApJ, 762, L21
MacGregor, M. A., Wilner, D. J., Andrews, S. M.,
Lestrade, J.-F., & Maddison, S. 2015, ApJ, 809,
47
MacGregor, M. A., Wilner, D. J., Chandler, C., et
2015, A&A, 576, A133
Malhotra, R. 1998, Solar System Formation and
Evolution, ASP Conference Series, 149, 37
Malo, L., Doyon, R., Lafreni`ere, D., et al. 2013,
ApJ, 762, 88
Marleau, G.-D., & Cumming, A. 2014, MNRAS,
437, 1378
Marley, M.S., Fortney, J.J., Hubickyj, O.,
Bodenheimer, P., Lissauer, J.J. 2007, ApJ, 655,
541
Marley, M. S., Saumon, D., Cushing, M., et al.
2012, ApJ, 754, 135
Marois, C., Macintosh, B., Barman, T., et al.
2008, Science, 322, 1348
Marois, C., Zuckerman, B., Konopacky, Q. M.,
Macintosh, B., & Barman, T. 2010, Nature,
468, 1080
Matthews, B., Kennedy, G., Sibthorpe, B., et al.
2014, ApJ, 780, 97
Mo´or, A., K´osp´al, ´A., ´Abrah´am, P., et al. 2015,
MNRAS, 447, 577
Mordasini, C. 2013, A&A, 558, A113
Mustill, A. J., & Wyatt, M. C. 2009, MNRAS,
399, 1403
Mustill, A. J., & Wyatt, M. C. 2012, MNRAS,
419, 3074
Pan, M., & Sari, R. 2005, Icarus, 173, 342
Pan, M., & Schlichting, H. E. 2012, ApJ, 747, 113
Patience, J., Bulger, J., King, R. R., et al. 2011,
A&A, 531, L17
Pearce, T. D., & Wyatt, M. C. 2014, MNRAS,
443, 2541
Pueyo, L., Soummer, R., Hoffmann, J., et al.
2015, ApJ, 803, 31
Quillen, A. C. 2006, MNRAS, 372, L14
Quillen, A. C., & Faber, P. 2006, MNRAS, 373,
1245
Regaly, Z., Dencs, Z., Moor, A., & Kovacs, T.
2017, arXiv:1710.01524
Reidemeister, M., Krivov, A. V., Schmidt,
T. O. B., et al. 2009, A&A, 503, 247
Ricci, L., Testi, L., Maddison, S. T., & Wilner,
D. J. 2012, A&A, 539, L6
Ricci, L., Carpenter, J. M., Fu, B., et al. 2015,
ApJ, 798, 124
Sadakane, K., & Nishida, M. 1986, PASP, 98, 685
Schuppler, C., Lohne, T., Krivov, A. V., et al.
al. 2016, ApJ, 823, 79
2015, A&A, 581, A97
18
Soummer, R., Brendan Hagan, J., Pueyo, L., et al.
Williams, J. P., & Andrews, S. M. 2006, ApJ, 653,
2011, ApJ, 741, 55
1480
Spiegel, D. S., & Burrows, A. 2012, ApJ, 745, 174
Su, K. Y. L., Rieke, G. H., Stapelfeldt, K. R., et
al. 2009, ApJ, 705, 314
Wilner, D. J., Andrews, S. M., & Hughes, A. M.
2011, ApJ, 727, L42
Wisdom, J. 1980, AJ, 85, 1122
Wright, D. J., Chen´e, A.-N., De Cat, P., et al.
Su, K. Y. L., MacGregor, M. A., Booth, M., et al.
2011, ApJ, 728, L20
2017, AJ, 154, 225
van Leeuwen, F. 2007, A&A, 474, 653
Wertz, O., Absil, O., G´omez Gonz´alez, C. A., et
al. 2017, A&A, 598, A83
Wyatt, M. C. 2006, ApJ, 639, 1153
Wyatt, M. C. 2008, ARAA, 46, 339
Zurlo, A., Vigan, A., Galicher, R., et al. 2016,
A&A, 587, A57
|
1611.09504 | 2 | 1611 | 2016-12-20T08:01:12 | A possible giant planet orbiting the cataclysmic variable LX Ser | [
"astro-ph.EP"
] | LX Ser is a deeply eclipsing cataclysmic variable with an orbital period of $0.^d 1584325$. Sixty two new eclipse times were determined by our observations and the AAVSO International Data base. Combining all available eclipse times, we analyzed the O-C behavior of LX Ser. We found that the O-C diagram of LX Ser shows a sinusoidal oscillation with a period of 22.8 yr and an amplitude of 0.00035 days. Two mechanisms (i.e., the Applegate mechanism and the light travel time effect) are applied to explain the cyclic modulation. We found that the Applegate mechanism is difficult to explain the cyclic oscillation in the orbital period. Therefore, the cyclic period change is most likely to be caused by the light travel time effect due to the presence of a third body. The mass of the tertiary component was determined to be $M_3\sim7.5 M_{Jup}$. We supposed that the tertiary companion is plausible a giant planet. The stability of the giant planet was checked, and we found that the multiple system is stable. | astro-ph.EP | astro-ph |
A possible giant planet orbiting the cataclysmic variable LX Ser
Li K.1,2, Hu, S.-M.1, Zhou, J.-L.3, Wu, D.-H.3, Guo, D.-F.1, Jiang, Y.-G.1, Gao, D.-Y.1,
Chen, X.1, Wang, X.-Y.1
ABSTRACT
LX Ser is a deeply eclipsing cataclysmic variable with an orbital period of
0.d1584325. Sixty two new eclipse times were determined by our observations
and the AAVSO International Data base. Combining all available eclipse times,
we analyzed the O − C behavior of LX Ser. We found that the O − C diagram of
LX Ser shows a sinusoidal oscillation with a period of 22.8 yr and an amplitude
of 0.00035 days. Two mechanisms (i.e., the Applegate mechanism and the light
travel time effect) are applied to explain the cyclic modulation. We found that
the Applegate mechanism is difficult to explain the cyclic oscillation in the orbital
period. Therefore, the cyclic period change is most likely to be caused by the
light travel time effect due to the presence of a third body. The mass of the
tertiary component was determined to be M3 ∼ 7.5 MJ up. We supposed that the
tertiary companion is plausible a giant planet. The stability of the giant planet
was checked, and we found that the multiple system is stable.
Subject headings: stars: binaries: eclipsing - stars: novae, cataclysmic variables
- stars: individual (LX Ser)
1.
Introduction
Cataclysmic variables (CVs) are interacting binary stars composed of a degenerate white
dwarf primary and a Roche lobe filling M dwarf secondary. LX Ser was first identified to be
an eclipsing CV by Stepanian (1979) when searching galaxies with ultraviolet continuum and
1Shandong Provincial Key Laboratory of Optical Astronomy and Solar-Terrestrial Environment, Institute
of Space Sciences, Shandong University, Weihai, 264209, China (e-mail: [email protected], [email protected]
(Li, K.), [email protected] (Hu, S.-M.))
2Key Laboratory for the Structure and Evolution of Celestial Objects, Chinese Academy of Sciences,
Kunming 650011, China
3School of Astronomy and Space Science and Key Laboratory of Modern Astronomy and Astrophysics in
Ministry of Education, Nanjing University, Nanjing 210093, China
– 2 –
was classified to be SW Sex subclass of CVs. Stepanian (1979) found that the brightness of
LX Ser was about mV = 14 mag. Recently, the orbital inclination and mass ratio of LX Ser
were determined to be i = 79◦.0 and q = 0.50 by Marin et al. (2007). By analyzing the eclipse
times of LX Ser, Horne (1980) determined an ephemeris HJD = 2444293.02377(±0.00020)+
0d.1584328(±0.0000010)E and showed that an upper limit on P is 10−5. Eason et al. (1984)
derived a new ephemeris using a second-order least-squares fitting of all the eclipse times,
and the second-order term was insignificant and was ignored. After that, many new eclipse
times were obtained (e.g., Agerer & Hubscher 2003; Diethelm 2003; Zejda 2004; Hubscher
et al. 2005; Krajci 2005, 2006; Zejda et al. 2006), allowing us to reanalyze the eclipse times.
Eclipsing times of CVs can be determined with high precision, and small amplitude
orbital period changes could be discovered. Therefore, CVs are very good targets to search
for planetary-mass companions. The presence of the planetary companions orbiting the
central stars will cause small wobbles of the barycenter of the triple system. The light from
the stars will travel closer or further due to the variation of the distance between the host
system and the earth. Then, the arrival eclipse times vary cyclically and the observed-
calculated (O − C) diagram shows a cyclic change. By analyzing the O − C variations, very
small mass companions can be detected. This method has been successfully used to detect
substellar companions surrounding CVs, e.g., Z Cha (Dai et al. 2009), DP Leo (Qian et
al. 2010a), QS Vir (Qian et al. 2010b), UZ For (Potter et al. 2011), V2051 Oph (Qian et
al. 2015). In this paper, we show the investigation of the eclipse times of LX Ser and the
possible giant planet orbiting this system.
2. New Observations
New photometric observations of LX Ser were carried out on April 24, May 30, and
June 26, 2015, and February 20, 2016. The CCD images were taken by using the 1.0 m
Cassegrain telescope at Weihai Observatory of Shandong University (Hu et al. 2014) with
the Andor DZ936 camera. The size of each pixel is 0.35′′, resulting that the effective field of
view is about 11.8′ × 11.8′. The observation information is listed in Table 1. All the CCD
images were analyzed using the IMRED and PHOT packages in IRAF1 procedure. The
light curve in Rc band observed on April 24, 2015 is displayed in Figure 1 for example. The
brightness flickering outside the eclipse can be seen, which is caused by variations of mass
transfer rate from the secondary red dwarf to the primary white dwarf. Four new times of
1IRAF is distributed by the National Optical Astronomy Observatories, which is operated by the Asso-
ciation of Universities for Research in Astronomy Inc., under contract to the National Science Foundation.
– 3 –
light minimum were determined and are shown in Table 2. Our new eclipse times are 3 or 4
significant figures, indicating very high precision. We also collected the data from AAVSO
(American Association of Variable Stars Observers) International Data base1. Based on the
data, 58 eclipse times were reanalyzed and are also listed in Table 2, the corresponding light
curves are displayed in Figure 2. All the eclipse times were determined by this work using
the parabolic fitting method and was converted to Barycentric Julian Dates (BJD) using the
software of Eastman et al. (2010). As seen in Figures 1 and 2, all the eclipse times can be
fitted by parabolic curves very well.
3.
Investigation of the O − C diagram
Since the discovery, LX Ser has been continuously monitored. Many highly precise
eclipse times were determined. All the eclipse times collected from literatures including our
62 new determined ones are listed in Table 3. The eclipse times collected form literatures
are checked again since some data were derived a long time ago and obtained by different
observatories or filters. Those having obviously erring are abandoned. Some of the eclipse
times were not given errors in the literatures. The errors of the visual data are assumed to
be 0.00100, while that of the photoelectric and CCD data are 0.00010. We constructed the
O − C diagram using the following linear ephemeris taken from O-C Gateway2 (the initial
epoch was converted to BJD),
Min.I = BJD2444293.02457 + 0d.1584325E.
(1)
The O − C values are listed in Table 3 and the corresponding O − C curve is plotted in the
upper panel of Figure 3. A linear function was used to fit the O − C values. By using the
1http://www.aavso.org/
2http://var2.astro.cz/ocgate/
Table 1: Observation information of LX Ser
Date
24 Apr., 2015
30 May, 2015
26 Jun., 2015
20 Feb., 2016
Duration time (hr) Filter Exposure time (s)
3.89
0.78
0.65
0.94
Rc
Rc
V
N1
100
100
100
30
1 N means no filter.
– 4 –
Rc
-0.5
0.0
0.5
m
1.0
Rc
1.5
2.0
m
-0.5
0.0
0.5
1.0
1.5
2.0
37.14 37.16 37.18 37.20 37.22 37.24 37.26 37.28 37.30 37.32
73.26
73.27
73.28
73.29
HJD2457100+
2457100+
-0.5
0.0
0.5
m
1.0
1.5
Rc
m
-0.5
0.0
0.5
1.0
1.5
2.0
N
0.01
0.02
0.03
0.04
2457200+
0.05
0.06
0.07
39.26
39.27
39.28
39.29
2457400+
Fig. 1.- The light curves of LX Ser observed using the 1.0 m Cassegrain telescope at
Weihai Observatory of Shandong University. The solid and open circles represent magnitude
differences between LX Ser and the comparison star and those between the comparison and
the check stars, while the solid lines refer to the parabolic fits for the eclipse times.
– 5 –
Table 2: New eclipse times of LX Ser
HJD
BJD
2452777.87530
2452778.82581
2452779.77639
2452779.93490
2452780.72718
2452780.88538
2452781.83613
2452782.78677
2452782.94532
2452783.73740
2452786.74781
2452786.90609
2452787.85687
2453500.48638
2453502.54540
2453506.50728
2453514.42838
2453516.48800
2453519.49811
2453521.39970
2453521.55769
2453541.52033
2454316.41428
2454580.52079
2454628.52582
2454976.44292
2454994.50413
2455001.47512
2455037.43966
2455662.45633
2455663.40660
2452777.87605
2452778.82656
2452779.77714
2452779.93565
2452780.72793
2452780.88613
2452781.83688
2452782.78752
2452782.94607
2452783.73815
2452786.74856
2452786.90684
2452787.85762
2453500.48711
2453502.54613
2453506.50801
2453514.42911
2453516.48873
2453519.49884
2453521.40043
2453521.55842
2453541.52106
2454316.41502
2454580.52153
2454628.52656
2454976.44368
2454994.50489
2455001.47588
2455037.44042
2455662.45711
2455663.40738
Errors
0.00008
0.00010
0.00008
0.00012
0.00015
0.00013
0.00012
0.00012
0.00013
0.00009
0.00013
0.00009
0.00010
0.00011
0.00031
0.00033
0.00037
0.00035
0.00042
0.00039
0.00040
0.00012
0.00008
0.00034
0.00013
0.00014
0.00006
0.00007
0.00017
0.00016
0.00022
HJD
Source
Cycle
53555 AAVSO 2455672.43733
53561 AAVSO 2455778.42896
53567 AAVSO 2456028.43535
53568 AAVSO 2456088.48114
53573 AAVSO 2456101.63073
53574 AAVSO 2456378.88725
53580 AAVSO 2456381.89715
53586 AAVSO 2456383.79856
53587 AAVSO 2456384.43209
53592 AAVSO 2456384.59069
53611 AAVSO 2456384.74908
53612 AAVSO 2456385.85843
53618 AAVSO 2456386.80920
58116 AAVSO 2456389.50214
58129 AAVSO 2456389.66136
58154 AAVSO 2456403.44384
58204 AAVSO 2456410.41514
58217 AAVSO 2456427.68469
58236 AAVSO 2456782.41510
58248 AAVSO 2456792.39590
58249 AAVSO 2456798.41636
58375 AAVSO 2457091.67550
63266 AAVSO 2457094.68559
64933 AAVSO 2457097.69575
65236 AAVSO 2457134.45164
67432 AAVSO 2457137.30350
67546 AAVSO 2457159.48411
67590 AAVSO 2457163.44506
67817 AAVSO 2457173.26766
71762 AAVSO 2457200.04314
71768 AAVSO 2457439.27618
BJD
2455672.43811
2455778.42974
2456028.43613
2456088.48192
2456101.63151
2456378.88804
2456381.89794
2456383.79935
2456384.43288
2456384.59148
2456384.74987
2456385.85922
2456386.80999
2456389.50293
2456389.66215
2456403.44463
2456410.41593
2456427.68548
2456782.41588
2456792.39668
2456798.41714
2457091.67627
2457094.68636
2457097.69652
2457134.45241
2457137.30427
2457159.48488
2457163.44583
2457173.26843
2457200.04391
2457439.27695
Errors
0.00016
0.00011
0.00010
0.00009
0.00012
0.00014
0.00019
0.00029
0.00015
0.00015
0.00016
0.00020
0.00017
0.00011
0.00021
0.00008
0.00016
0.00020
0.00006
0.00020
0.00006
0.00007
0.00021
0.00005
0.00004
0.00007
0.00005
0.00013
0.00015
0.00049
0.00004
Cycle
Source
71825 AAVSO
72494 AAVSO
74072 AAVSO
74451 AAVSO
74534 AAVSO
76284 AAVSO
76303 AAVSO
76315 AAVSO
76319 AAVSO
76320 AAVSO
76321 AAVSO
76328 AAVSO
76334 AAVSO
76351 AAVSO
76352 AAVSO
76439 AAVSO
76483 AAVSO
76592 AAVSO
78831 AAVSO
78894 AAVSO
78932 AAVSO
80783 AAVSO
80802 AAVSO
80821 AAVSO
81053 AAVSO
81071
81211 AAVSO
81236 AAVSO
81298
81467
82977
1m
1m
1m
1m
– 6 –
13.5
14.0
14.5
15.0
m
15.5
m
0.75
0.80
0.85
0.90
0.70
0.75
16.0
16.5
17.0
2452777+
0.80
2452778+
0.85
0.90
0.95
14.0
14.5
15.0
15.5
14.0
14.5
15.0
15.5
16.0
16.5
17.0
14.0
14.5
15.0
15.5
13.5
14.0
14.5
15.0
m
15.5
0.75
0.80
0.85
2452780+
0.90
0.95
16.0
16.5
17.0
14.0
14.5
15.0
15.5
0.75
0.80
0.85
2452779+
0.90
0.95
m
16.0
m
16.0
m
16.0
0.72
0.81
2452781+
0.90
0.75
0.80
0.85
0.90
0.95
2452787+
0.38
0.40
0.42
2453514+
0.44
0.46
m
m
16.5
17.0
17.5
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
14.0
14.5
15.0
0.70
0.75
0.80
0.85
0.90
0.95
0.70
0.72
0.74
0.76
16.5
17.0
17.5
2452782+
0.78
0.80
2452783+
0.82
0.84
0.86
0.88
14.0
14.5
15.0
15.5
0.75
0.80
0.85
2452786+
0.90
0.95
16.5
17.0
17.5
14.5
15.0
15.5
16.0
m
16.0
m
16.5
0.40
0.45
0.50
2453500+
0.55
0.60
16.5
17.0
17.5
18.0
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
18.5
m
0.45
0.50
0.55
0.60
2453502+
0.42
0.44
0.46
0.48
2453516+
0.50
0.52
0.54
0.40
0.45
0.50
0.55
2453519+
14.0
14.5
15.0
15.5
m
17.0
17.5
18.0
18.5
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
18.5
14.0
14.5
15.0
15.5
0.40
0.45
0.50
2453506+
0.55
0.60
0.40
0.45
0.50
2453521+
0.55
m
15.5
16.0
m
m
16.0
16.0
16.5
17.0
16.5
17.0
17.5
0.54
0.56
0.58
0.38
0.39
0.40
0.41
0.42
0.43
0.44
0.49
0.50
0.51
0.44
0.46
0.48
0.50
0.52
2453541+
2454316+
15.0
15.5
16.0
0.52
0.53
2454580+
0.54
0.55
0.56
0.49
0.50
0.51
0.52
2454628+
0.53
0.54
0.55
16.5
17.0
17.5
14.5
15.0
15.5
16.0
m
16.5
m
16.5
17.0
17.5
17.0
17.5
18.0
m
14.5
15.0
15.5
16.0
16.5
17.0
17.5
13.5
14.0
14.5
15.0
m
15.5
16.0
16.5
17.0
m
14
15
16
17
14
15
m
16
17
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
18.5
m
14.5
15.0
15.5
16.0
m
16.5
17.0
17.5
18.0
14.0
14.5
15.0
15.5
16.0
16.5
17.0
m
0.42
0.43
0.44
0.45
0.50
0.51
0.52
0.465
0.470
0.475
0.480
2454976+
2454994+
2455001+
0.43
0.44
2455037+
0.45
Fig. 2.- The light curves of LX Ser using the AAVSO data. The solid circles represent the
visual magnitude of LX Ser, while the solid lines refer to the parabolic fits for the eclipse
times.
– 7 –
14.5
15.0
15.5
16.0
15.0
15.5
16.0
14.0
14.5
15.0
15.5
m
16.5
m
16.5
m
16.0
0.445
0.450
0.455
2455662+
0.460
0.42
0.43
0.44
2456028+
m
17.0
17.5
18.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
14.5
15.0
15.5
16.0
0.39
0.40
0.41
0.42
2455663+
0.47
0.48
2456088+
0.49
m
16.5
m
17.0
17.5
18.0
0.75
0.80
0.85
2456381+
0.90
0.95
0.80
0.85
0.90
2456382+
0.95
0.430
0.435
2455672+
0.440
17.0
17.5
14.5
15.0
15.5
16.0
16.5
17.0
17.5
14.5
15.0
15.5
16.0
0.41
0.42
0.43
0.44
2455778+
m
16.5
m
16.5
0.62
0.64
0.66
0.70
0.68
2456101+
17.0
17.5
18.0
18.5
0.72
0.74
0.76
0.82
0.84
0.86
0.88
0.90
0.92
0.94
0.96
0.98
1.00
2456378+
m
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
0.70
0.75
0.80
2456383+
0.85
0.90
0.40
0.45
0.50
0.55
0.60
0.65
2456384+
0.70
0.75
0.80
0.85
14.5
15.0
15.5
16.0
m
16.5
0.85
0.90
0.45
0.50
0.55
0.60
0.65
0.81
0.84
0.87
0.90
0.93
2456385+
m
m
14.5
15.0
15.5
16.0
16.5
17.0
17.5
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
0.70
0.75
0.80
2456386+
17.0
17.5
18.0
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
m
m
14.5
15.0
15.5
16.0
16.5
17.0
15.0
15.5
16.0
16.5
17.0
17.5
18.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
14.5
15.0
15.5
16.0
16.5
17.0
14.0
14.5
15.0
15.5
m
m
m
m
16.0
16.5
17.0
17.5
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
m
m
14.5
15.0
15.5
16.0
m
16.5
17.0
17.5
18.0
17.0
17.5
18.0
14.5
15.0
15.5
0.43
0.44
2456403+
0.45
16.0
m
16.5
17.0
17.5
0.42
0.43
0.38
0.39
0.40
0.41
2456792+
m
14.5
15.0
15.5
16.0
16.5
17.0
17.5
0.69
0.70
0.71
0.64
0.65
0.66
0.67
0.68
2457097+
0.69
0.70
0.71
0.72
2456389+
0.41
2456782+
0.67
0.68
2457094+
0.39
0.40
0.41
2456410+
0.42
0.43
0.60
0.68
0.72
0.40
0.64
2456427+
0.40
0.41
2456798+
0.42
m
m
14.5
15.0
15.5
16.0
16.5
17.0
17.5
14.5
15.0
15.5
16.0
16.5
17.0
17.5
m
14.0
14.5
15.0
15.5
16.0
16.5
17.0
17.5
18.0
0.65
0.66
0.67
0.69
0.68
2457091+
0.70
0.71
0.64
0.65
0.66
14.5
15.0
15.5
16.0
m
16.5
17.0
17.5
0.43
0.44
0.45
2457134+
0.46
0.47
0.45
0.46
0.47
0.48
0.49
0.50
0.43
0.44
0.45
0.46
2457159+
2457163+
Fig. 2.- - continued
– 8 –
weighted least-squares method, we determined the following equation
Min.I = BJD2444293.024192(±0.000083) + 0d.158432491(±0.000000002) × E.
(2)
The variance is σ1 = 5.14 × 10−7 and the fitting curve is displayed in Figure 3. During the
fitting, the weight for the visual data is 1, while that for the photoelectric and CCD data
is 10. As seen in the lower panel of Figure 3, a possible small amplitude oscillation can be
extracted. Therefore, we intended to use a simultaneous linear-plus-sinusoidal function to fit
the O − C curve. In order to determine the period of the sinusoidal variation, the Period04
package (Lenz & Breger 2005) was employed. The resulting Fourier spectrum is shown in
Figure 4. A peak at f = 1.90386493(±0.06022577) × 10−5 was obtained, meaning that the
period of the sinusoidal variation is 22.80(±0.72) yr.
Using the weighted least-squares method, we determined the following equation,
Min.I = BJD2444293.023557(±0.000191) + 0d.158432495(±0.000000002) × E +
+0d.00035(±0.00008) × sin [0◦.00685 × E + 32◦.8(±0.3)].
(3)
The variance is σ2 = 4.67 ×10−7, which is smaller than the value of the linear fit. The F -test
method as discussed by Pringle (1975) was used to estimate to what extent the statistical
significance is improved between the linear fit and the linear-plus-sinusoidal fit. The statistic
parameter λ is corrected to be
λ =
1 − σ2
(σ2
2)/2
σ2
2/(n − 4)
.
(4)
In this equation, n is the number of eclipse times. We obtained F (2, 184) = 19.5, indicating
that the statistical significance of the linear-plus-sinusoidal fit with respect to the linear fit
is more than the 99.99% level. The best fitting curve for the linear-plus-sinusoidal terms is
displayed in the upper panel of Figure 5. The linear term of Equation (3) represents the
revision on the initial linear ephemeris (Equation 1). When the linear term was removed, the
(O − C)2 values are displayed in the middle panel of Figure 5. A 22.8 yr cyclic modulation
with an amplitude of 0.00035 days can be seen. The residuals from Equation (3) are shown
in the lower panel of Figure 5, where no regular changes can be subtracted.
4. Discussion
4.1. The cyclic orbital period variation
Based on the investigation of the O − C diagram, we found that the orbital period
of LX Ser shows a cyclic modulation with a period of P3 = 22.8 yr and an amplitude of
– 9 –
vis
pe&CCD
AAVSO data
our observations
0
20000
40000
E
60000
80000
100000
0.002
0.000
)
-
C
O
(
-0.002
-0.004
0.004
0.002
0.000
-0.002
-0.004
l
s
a
u
d
s
e
R
i
Fig. 3.- The O − C diagram of LX Ser. Upper panel shows the linear fit of the O − C
curve. Lower panel displays the residuals. "Vis" represents the visual data, "pe" refers to
the photoelectric data, and "CCD" displays the CCD data.
1.90386493e-005
e
d
u
t
i
l
p
m
A
0.0005
0.0004
0.0003
0.0002
0.0001
0.0000
0.00000
0.00005
0.00010
E-1
0.00015
0.00020
Fig. 4.- The fourier spectrum for the orbital period variation.
– 10 –
Table 3: All avaiable eclipse times of LX Ser
Cycle
Errors Method
HJD
2400000+
44293.02430
44312.98580
44316.94620
44320.90690
44343.72270
44344.83060
44346.89130
44372.71520
44384.75580
44384.91420
44385.86550
44396.79650
44644.58600
44691.48200
44691.64200
44731.40600
44755.48800
44755.48900
44808.40600
44809.35400
44809.35500
44811.41500
44812.36500
44812.36600
44817.43700
44848.33000
44869.24300
44987.59200
44994.72100
45053.49700
45061.57900
45101.50500
45104.51400
45114.49600
45115.44700
45115.44700
45139.52900
45142.53800
45148.39900
45208.28600
45358.64100
45385.57400
45432.46900
45432.47000
45439.44100
45442.44900
45459.40400
45548.44300
45548.44300
45701.64500
45741.57200
45766.60400
45766.60500
45879.40700
45906.34200
45908.40100
45911.41100
46046.71400
46154.60600
BJD
2400000+
44293.02487
44312.98637
44316.94677
44320.90747
44343.72327
44344.83117
44346.89187
44372.71576
44384.75636
44384.91476
44385.86606
44396.79706
44644.58656
44691.48255
44691.64255
44731.40655
44755.48855
44755.48955
44808.40656
44809.35456
44809.35556
44811.41556
44812.36556
44812.36656
44817.43756
44848.33056
44869.24356
44987.59256
44994.72156
45053.49756
45061.57956
45101.50556
45104.51456
45114.49656
45115.44756
45115.44756
45139.52956
45142.53856
45148.39956
45208.28657
45358.64157
45385.57457
45432.46957
45432.47057
45439.44157
45442.44957
45459.40457
45548.44358
45548.44358
45701.64559
45741.57259
45766.60459
45766.60559
45879.40759
45906.34259
45908.40159
45911.41159
46046.71460
46154.60660
0.00030
–
–
–
–
–
–
–
0.00050
0.00030
0.00030
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
0
126
151
176
320
327
340
503
579
580
586
655
2219
2515
2516
2767
2919
2919
3253
3259
3259
3272
3278
3278
3310
3505
3637
4384
4429
4800
4851
5103
5122
5185
5191
5191
5343
5362
5399
5777
6726
6896
7192
7192
7236
7255
7362
7924
7924
8891
9143
9301
9301
10013
10183
10196
10215
11069
11750
O-C
(O-C)a
2
(O-C)b
3
References
0
-0.00100
-0.00141
-0.00152
0.00000
-0.00113
-0.00005
-0.00065
-0.00092
-0.00095
-0.00025
-0.00109
-0.00003
-0.00005
0.00152
-0.00104
-0.00078
0.00022
0.00077
-0.00182
-0.00082
-0.00044
-0.00104
-0.00004
0.00112
-0.00022
-0.00031
-0.00039
-0.00085
-0.00331
-0.00137
-0.00036
-0.00157
-0.00082
-0.00042
-0.00042
-0.00016
-0.00137
-0.00238
-0.00285
-0.00029
-0.00082
-0.00184
-0.00084
-0.00087
-0.00308
-0.00036
-0.00041
-0.00041
-0.00264
-0.00063
-0.00096
0.00004
-0.00190
-0.00042
-0.00105
-0.00126
0.00039
-0.00014
0.00067
-0.00031
-0.00072
-0.00083
0.00068
-0.00044
0.00063
0.00003
-0.00023
-0.00026
0.00043
-0.00040
0.00067
0.00065
0.00222
-0.00033
-0.00007
0.00092
0.00148
-0.00110
-0.00010
0.00026
-0.00032
0.00067
0.00183
0.00049
0.00040
0.00033
-0.00012
-0.00258
-0.00064
0.00037
-0.00084
-0.00009
0.00031
0.00031
0.00057
-0.00063
-0.00164
-0.00211
0.00045
-0.00007
-0.00109
-0.00008
-0.00011
-0.00233
0.00038
0.00034
0.00034
-0.00187
0.00013
-0.00019
0.00080
-0.00112
0.00035
-0.00027
-0.00048
0.00117
0.00064
0.00082
Horne 1980
-0.00017 Africano & Klimke 1981
-0.00058 Africano & Klimke 1981
-0.00069 Africano & Klimke 1981
0.00081 Africano & Klimke 1981
-0.00031 Africano & Klimke 1981
0.00076 Africano & Klimke 1981
0.00016 Africano & Klimke 1981
Young et al. 1981
-0.00011
Young et al. 1981
-0.00014
0.00056
Young et al. 1981
-0.00028 Africano & Klimke 1981
0.00073
BBSAG 53
BBSAG 53
0.00070
BBSAG 53
0.00228
BBSAG 54
-0.00028
BBSAG 54
-0.00002
0.00097
BBSAG 54
BBSAG 56
0.00151
BBSAG 56
-0.00107
BBSAG 56
-0.00007
BBSAG 56
0.00030
-0.00029
BBSAG 56
BBSAG 56
0.00070
BBSAG 56
0.00186
BBSAG 56
0.00052
0.00043
BBSAG 56
BBSAG 58
0.00033
BBSAG 58
-0.00012
BBSAG 59
-0.00258
BBSAG 60
-0.00064
0.00036
BBSAG 61
BBSAG 60
-0.00085
BBSAG 61
-0.00010
BBSAG 60
0.00030
BBSAG 60
0.00030
0.00055
BBSAG 61
BBSAG 61
-0.00065
BBSAG 61
-0.00166
BBSAG 62
-0.00213
BBSAG 64
0.00041
-0.00011
BBSAG 65
BBSAG 66
-0.00113
BBSAG 66
-0.00013
BBSAG 66
-0.00016
BBSAG 66
-0.00237
0.00034
BBSAG 66
BBSAG 68
0.00029
BBSAG 68
0.00029
BBSAG 70
-0.00192
0.00008
BBSAG 71
BBSAG 71
-0.00024
BBSAG 71
0.00075
BBSAG 72
-0.00117
BBSAG 73
0.00030
-0.00032
BBSAG 73
BBSAG 73
-0.00053
BBSAG 75
0.00112
0.00061
BBSAG 76
– 11 –
Table 3: - continued
HJD
2400000+
46175.51800
46183.44000
46186.45000
46252.35800
46296.40200
46497.61200
46497.61300
46622.45600
46625.46600
46831.58700
47003.32900
47023.44700
47212.61700
47304.50800
47383.40700
47535.66200
47746.37600
47890.70900
47942.51500
48039.47600
48306.59300
49004.64500
49158.48300
49475.50700
49799.50400
49836.57600
50139.65700
50539.54100
50864.48600
50988.37900
51603.57310
51641.43900
52072.53400
52320.63969
52348.52400
52410.47070
52730.50400
52777.87530
52778.82581
52779.77639
52779.93490
52780.72718
52780.88538
52781.83613
52782.78677
52782.94532
52783.73740
52786.74781
52786.90609
52787.85687
52828.41650
53146.23100
53436.63800
53465.63130
53498.26840
53500.48638
53500.80300
53500.96130
BJD
2400000+
46175.51860
46183.44060
46186.45060
46252.35862
46296.40262
46497.61262
46497.61362
46622.45663
46625.46663
46831.58764
47003.32964
47023.44764
47212.61765
47304.50866
47383.40766
47535.66266
47746.37665
47890.70965
47942.51566
48039.47666
48306.59367
49004.64566
49158.48366
49475.50767
49799.50469
49836.57669
50139.65771
50539.54172
50864.48675
50988.37975
51603.57388
51641.43978
52072.53477
52320.64046
52348.52476
52410.47146
52730.50475
52777.87605
52778.82656
52779.77714
52779.93565
52780.72793
52780.88613
52781.83688
52782.78752
52782.94607
52783.73815
52786.74856
52786.90684
52787.85762
52828.41725
53146.23174
53436.63873
53465.63203
53498.26913
53500.48711
53500.80373
53500.96203
Errors Method
Cycle
O-C
(O-C)a
2
(O-C)b
3
References
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
0.00100
0.00200
0.00100
–
0.00200
0.00100
0.00200
0.00100
0.00300
0.00200
–
0.00090
–
0.00090
–
0.00020
0.00100
0.00008
0.00010
0.00008
0.00012
0.00015
0.00013
0.00012
0.00012
0.00013
0.00009
0.00013
0.00009
0.00010
0.00090
0.00030
0.00200
0.00030
–
0.00011
0.00020
0.00030
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
vis
ccd
vis
ccd
vis
pe
vis
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
pe
ccd
ccd
vis
ccd
ccd
pe
ccd
ccd
11882
11932
11951
12367
12645
13915
13915
14703
14722
16023
17107
17234
18428
19008
19506
20467
21797
22708
23035
23647
25333
29739
30710
32711
34756
34990
36903
39427
41478
42260
46143
46382
49103
50669
50845
51236
53256
53555
53561
53567
53568
53573
53574
53580
53586
53587
53592
53611
53612
53618
53874
55880
57713
57896
58102
58116
58118
58119
-0.00123
-0.00086
-0.00107
-0.00098
-0.00121
-0.00048
0.00052
-0.00129
-0.00150
-0.00118
-0.00001
-0.00293
-0.00132
-0.00117
-0.00156
-0.00019
-0.00142
-0.00042
-0.00184
-0.00153
-0.00172
-0.00332
-0.00328
-0.00270
-0.00015
-0.00135
-0.00171
-0.00132
-0.00136
-0.00257
-0.00184
-0.00131
-0.00114
-0.00075
-0.00057
-0.00097
-0.00134
-0.00135
-0.00144
-0.00145
-0.00138
-0.00126
-0.00149
-0.00134
-0.00129
-0.00117
-0.00126
-0.00106
-0.00122
-0.00103
-0.00012
-0.00123
-0.00101
-0.00086
-0.00085
-0.00093
-0.00118
-0.00131
-0.00043
-0.00006
-0.00027
-0.00018
-0.00041
0.00032
0.00132
-0.00047
-0.00068
-0.00034
0.00083
-0.00208
-0.00046
-0.00031
-0.00069
0.00068
-0.00053
0.00046
-0.00094
-0.00062
-0.00080
-0.00236
-0.00231
-0.00171
0.00085
-0.00034
-0.00068
-0.00026
-0.00028
-0.00149
-0.00072
-0.00019
0
0.00040
0.00059
0.00018
-0.00015
-0.00016
-0.00025
-0.00026
-0.00019
-0.00007
-0.00030
-0.00015
-0.00010
0.00001
-0.00007
0.00012
-0.00003
0.00015
0.00106
-0.00002
0.00021
0.00036
0.00037
0.00029
0.00004
-0.00008
BBSAG 76
-0.00047
BBSAG 77
-0.00009
BBSAG 77
-0.00031
BBSAG 77
-0.00021
BBSAG 78
-0.00043
BBSAG 79
0.00033
BBSAG 79
0.00133
BBSAG 80
-0.00045
BBSAG 81
-0.00066
BBSAG 82
-0.00029
BBSAG 84
0.00092
BBSAG 85
-0.00199
BBSAG 87
-0.00033
BBSAG 88
-0.00015
BBSAG 89
-0.00052
BBSAG 90
0.00089
BBSAG 92
-0.00027
BBSAG 93
0.00076
BBSAG 94
-0.00063
BBSAG 95
-0.00029
BBSAG 97
-0.00041
BBSAG 103
-0.00185
BBSAG 104
-0.00178
BBSAG 106
-0.00116
BBSAG 108
0.00140
BBSAG 109
0.00020
BBSAG 111
-0.00015
BBSAG 114
0.00020
BBSAG 117
0.00011
BBSAG 118
-0.00111
BRNO 32
-0.00051
OEJV 74
0.00001
BBSAG 125
0.00007
OEJV 74
0.00040
0.00059
BBSAG 127
0.00017 Agerer & Hubscher 2003
-0.00025
Diethelm 2003
This paper
-0.00027
This paper
-0.00036
This paper
-0.00037
This paper
-0.00030
-0.00018
This paper
This paper
-0.00041
This paper
-0.00026
This paper
-0.00021
This paper
-0.00009
-0.00018
This paper
This paper
0.00001
This paper
-0.00014
This paper
0.00004
Zejda 2004
0.00094
-0.00020
Krajci 2005
OEJV 3
-0.00001
Zejda et al. 2006
0.00013
VSOLJ 44
0.00013
0.00005
This paper
Krajci 2006
-0.00019
-0.00032
Krajci 2006
– 12 –
Table 3: - continued
HJD
2400000+
53501.75400
53501.91160
53502.54540
53504.76360
53504.92380
53506.50728
53510.46830
53514.42838
53516.48800
53519.49811
53521.39970
53521.55769
53541.52033
54218.34361
54316.41428
54580.52079
54628.52582
54976.44292
54994.50413
55001.47512
55037.43966
55662.45633
55663.40660
55672.43733
55778.42896
55988.66900
56028.43535
56088.48114
56101.63073
56378.88725
56381.89715
56383.79856
56384.43209
56384.59069
56384.74908
56385.85843
56386.80920
56389.50214
56389.66136
56403.44384
56410.41514
56427.68469
56782.41510
56792.39590
56798.41636
57091.67550
57094.68559
57097.69575
57132.39190
57134.45164
57135.40230
57137.30350
57159.48411
57163.44506
57173.26766
57200.04314
57439.27618
BJD
2400000+
53501.75473
53501.91233
53502.54613
53504.76433
53504.92453
53506.50801
53510.46903
53514.42911
53516.48873
53519.49884
53521.40043
53521.55842
53541.52106
54218.34434
54316.41502
54580.52153
54628.52656
54976.44368
54994.50489
55001.47588
55037.44042
55662.45711
55663.40738
55672.43811
55778.42974
55988.66978
56028.43613
56088.48192
56101.63151
56378.88804
56381.89794
56383.79935
56384.43288
56384.59148
56384.74987
56385.85922
56386.80999
56389.50293
56389.66215
56403.44463
56410.41593
56427.68548
56782.41588
56792.39668
56798.41714
57091.67627
57094.68636
57097.69652
57132.39267
57134.45241
57135.40307
57137.30427
57159.48488
57163.44583
57173.26843
57200.04391
57439.27695
Errors Method
Cycle
O-C
(O-C)a
2
(O-C)b
3
References
0.00060
0.00020
0.00031
0.00020
0.00090
0.00033
0.00050
0.00037
0.00035
0.00042
0.00039
0.00040
0.00012
0.00020
0.00008
0.00034
0.00013
0.00014
0.00006
0.00007
0.00017
0.00016
0.00022
0.00016
0.00011
0.00100
0.00010
0.00009
0.00012
0.00014
0.00019
0.00029
0.00015
0.00015
0.00016
0.00020
0.00017
0.00011
0.00021
0.00008
0.00016
0.00020
0.00006
0.00020
0.00006
0.00007
0.00021
0.00005
0.00030
0.00004
0.00020
0.00007
0.00005
0.00013
0.00015
0.00049
0.00004
ccd
ccd
pe
ccd
ccd
pe
ccd
pe
pe
pe
pe
pe
pe
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
ccd
pe
ccd
pe
ccd
ccd
ccd
ccd
ccd
ccd
58124
58125
58129
58143
58144
58154
58179
58204
58217
58236
58248
58249
58375
62647
63266
64933
65236
67432
67546
67590
67817
71762
71768
71825
72494
73821
74072
74451
74534
76284
76303
76315
76319
76320
76321
76328
76334
76351
76352
76439
76483
76592
78831
78894
78932
80783
80802
80821
81040
81053
81059
81071
81211
81236
81298
81467
82977
-0.00077
-0.00160
-0.00153
-0.00139
0.00038
-0.00046
-0.00026
-0.00099
-0.00099
-0.00110
-0.00070
-0.00114
-0.00100
-0.00135
-0.00040
-0.00086
-0.00088
-0.00153
-0.00162
-0.00166
-0.00130
-0.00082
-0.00115
-0.00107
-0.00078
-0.00067
-0.00088
-0.00100
-0.00131
-0.00166
-0.00198
-0.00176
-0.00196
-0.00179
-0.00183
-0.00151
-0.00133
-0.00175
-0.00096
-0.00211
-0.00184
-0.00143
-0.00140
-0.00184
-0.00182
-0.00125
-0.00137
-0.00143
-0.00200
-0.00188
-0.00182
-0.00181
-0.00175
-0.00161
-0.00183
-0.00144
-0.00147
0.00045
-0.00037
-0.00030
-0.00016
0.00160
0.00076
0.00096
0.00023
0.00023
0.00012
0.00052
0.00008
0.00023
-0.00008
0.00087
0.00043
0.00041
-0.00021
-0.00030
-0.00034
0.00001
0.00053
0.00020
0.00028
0.00058
0.00070
0.00050
0.00038
0.00007
-0.00026
-0.00058
-0.00036
-0.00056
-0.00039
-0.00043
-0.00011
0.00006
-0.00035
0.00043
-0.00071
-0.00044
-0.00002
0.00002
-0.00041
-0.00039
0.00018
0.00007
0.00001
-0.00055
-0.00043
-0.00037
-0.00036
-0.00030
-0.00016
-0.00038
0
0
Krajci 2006
0.00021
Krajci 2006
-0.00061
This paper
-0.00054
Krajci 2006
-0.00040
Krajci 2006
0.00136
0.00052
This paper
0.00073 Hubscher et al. 2005
0.00000
This paper
This paper
0.00000
This paper
-0.00011
This paper
0.00028
This paper
-0.00015
-0.00001
This paper
OEJV 107
-0.00034
This paper
0.00061
This paper
0.00018
This paper
0.00017
-0.00040
This paper
This paper
-0.00049
This paper
-0.00053
This paper
-0.00016
This paper
0.00047
0.00015
This paper
This paper
0.00023
This paper
0.00055
OEJV 147
0.00072
This paper
0.00053
0.00042
This paper
This paper
0.00012
This paper
-0.00014
This paper
-0.00046
-0.00024
This paper
This paper
-0.00044
This paper
-0.00027
This paper
-0.00031
This paper
0.00000
0.00018
This paper
This paper
-0.00023
This paper
0.00055
This paper
-0.00058
This paper
-0.00031
0.00009
This paper
This paper
0.00022
This paper
-0.00021
This paper
-0.00019
This paper
0.00044
0.00033
This paper
This paper
0.00027
Hubscher 2016
-0.00029
This paper
-0.00017
Hubscher 2016
-0.00011
-0.00010
This paper
This paper
-0.00003
This paper
0.00010
This paper
-0.00011
0.00028
This paper
This paper
0.00029
a (O-C)2 means the residuals for Equation (2), b (O-C)3 means the residuals for Equation (3)
– 13 –
A3 = 0.00035 days. The cyclic change of the O − C diagram can be explained either by
the Applegate mechanism (Applegate 1992) due to the magnetic activity of the red dwarf
component or by the light travel time effect due to a tertiary companion.
The Applegate mechanism was proposed according to the conclusion determined by
Hall (1989) who analyzed the orbital period variations of 101 Algols. Hall (1989) found that
all the binary systems which show cyclic period changes have late spectral type secondary
components. However, a recently statistical study by Liao & Qian (2010) reveals that the
percentages for both late type and early type binary systems that exhibit cyclic orbital period
changes are similar, and they concluded that the cyclic oscillation is most likely caused by
the light travel time effect via the presence of a third body.
According to Applegate mechanism, solar-like magnetic cycles would result in shape
changes of the low-mass components, thus redistributing the angular momentum within
the interior of the star. Then, the oblateness is changed, causing a change of the stellar
quadrupole moment which consequently leads to the variation of orbital period. Using the
equation
∆P
P
O − C
Pmod
= 2π
,
(5)
where (O − C) and Pmod are the amplitude and modulation period of the cyclic oscillation
of the orbital period. The fractional period change ∆P
P was determined to be 2.67 × 10−7.
By considering a typical mass of M1 = 0.6 M⊙ for the primary white dwarf of LX Ser, the
mass of the secondary red dwarf can be calculated to be M2 = q × M1 =0.3 M⊙ (the mass
ratio q is 0.5 based on the result of Marin et al. 2007). The radius of the secondary red
dwarf was estimated to be 0.38 R⊙ based on Cox (2000), the binary separation a = 1.19R⊙
can be derived using M1 + M2 = 0.0134a3/P 2. Then, according to the equation
∆P
P
= −9(
R2
a
)2 ∆Q
M2R2
2
,
(6)
we determined the variation of the quadrupole moment of the red dwarf to be ∆Q = −1.21 ×
1047 g cm2. The secondary component of LX Ser may be a full convective star, we can use
the same method that Brinkworth et al. (2006) used to calculate the energy required to
produce the cyclic modulation of LX Ser. We split the whole star into an inner core (denotes
as 1) and an outer shell (denotes as 2). By employing the Lane-Emden equation for an
n = 1.5 polytrope, different shell mass versus the required energy ∆E can be calculated
and is displayed in Figure 6. The minimum value of the required energy is achieved to be
∆Emin = 8.95 × 1040 erg when Ms = 0.224M⊙. Assuming that the temperature of T2 = 3500
K, the total radiated energy of the secondary red dwarf over the whole modulation period
is also shown in Figure 6. Obviously, the Applegate mechanism has difficulty to explain the
cyclic oscillation in the O − C diagram.
– 14 –
vis
pe&CCD
AAVSO data
our observations
0
20000
40000
E
60000
80000
100000
0.002
0.000
-0.002
-0.004
-0.006
0.004
0.002
0.000
-0.002
-0.004
0.004
0.002
0.000
-0.002
-0.004
1
)
-
C
O
(
2
)
-
C
O
(
.
s
e
R
Fig. 5.- The best fitting curve for the linear-plus-sinusoidal terms of LX Ser.
Required energy of Applegate mechanism
Total radiated energy of the secondary in 22.8 years
44
43
42
41
)
s
g
r
e
(
E
g
o
L
0.05
0.10
0.15
Ms(Msun)
0.20
0.25
0.30
Fig. 6.- The required energy ∆E versus different shell mass. The lowest value of required
energy is 8.95 × 1040 erg at a shell-mass of 0.224 M⊙.
– 15 –
Therefore, liking many other CVs, DQ Her (Dai & Qian 2009), UZ For (Potter et al.
2011), AM Her (Dai et al. 2013), the plausible explanation of the cyclic modulation is the
light travel time effect. Using the equation
f (m) =
(m3 sin i)3
(m1 + m2 + m3)2 =
4π
GP 2
3
× (a12 sin i)3,
(7)
the mass function f (m) = 4.29(±2.94) × 10−7 M⊙ and the mass of the third companion
M3 sin i′ = 0.0071(±0.0027) M⊙ = 7.4(±2.8) MJ up were computed. Simulations by Bonnell
& Bate (1994) and Holman & Wiegert (1999) exhibit that circumbinary planets have stable
orbits and are most possibly coplanar with the central eclipsing host systems. Considering
the orbital inclination of LX Ser i = 79◦.0, the mass of the third body is derived to be
M3 = 0.0072(±0.0027) M⊙ = 7.5(±2.8) MJ up with a separation of 9.12(±4.03) AU. Based
on the theoretical analysis (e.g., Chabrier & Baraffe 2000; Burrows et al. 2001), the upper
limit mass of a planet is about 0.0140 M⊙ ∼ 14.6 MJ up, the third companion is most likely
to be a giant planet.
4.2. The stability of the possible planet
In this subsection, we discussed the stability of the possible planet. As the mass of LX
Ser is much larger than the outer planet, we can ignore it when considering the evolution of
the inner CV. The logarithmically differentiating of semi-major axis of LX Ser with respect
to time gives
a
a
=
2 Jorb
Jorb
+
2(− M2)
M2
(1 −
M2
M1
).
(8)
When assuming a constant angular momentum, the period of the orbit increases, which may
impact the stability of the outer planet. However, there are many factors that influence
the orbital angular momentum in close binary systems, especially in a CV system. The
dominant angular momentum losses (AML) mechanism in long-period systems(Porb ≥ 3hrs)
is "magnetic braking", whereas short period CVs(Porb ≤ 2hrs) are assumed to be driven
by AML associated with the emission of "gravitational radiation" (Knigee & Baraffe 2011).
Apart from the two major factors, we should also consider mass transfer, mass accretion,
common-envelope evolution, wind accretion, tidal evolution and many other factors (Hurley
& Tout 2002). These factors are important in the orbital evolution of CVs. We adopted
the algorithm illustrated in Hurley & Tout (2002) and did a simulation with M1 = 0.6M⊙,
M2 = 0.3M⊙, P = 0.d158, e = 0, z = 0.02, where e is the eccentricity and z is the metallicity.
The evolutions of the masses of the two components, orbital period of LX Ser are shown in
Figure 7.
– 16 –
With multiple factors considered, we can see that the mass changes of the two compo-
nents occur at different time and speed, which is mainly caused by wind accretion. Besides,
the orbit of LX Ser contracts at the beginning of the evolution, mainly caused by gravita-
tional mechanism. With the evolution of the secondary component, the period of the system
is growing at a slow speed. Even at the universe timescale, the period is at most 0.d16, with-
out much difference with its initial period. Therefore, while considering the orbital evolution
of the outer planet, we can regard the inner CV as a particle with decreasing mass. The
outer planet will drift slowly away from the inner CV, the system remains stable.
Now we consider whether the existence of a planet with the characteristics derived in
the previous subsection is compatible with the evolution of LX Ser from a wide orbit binary
to a CV. Similar consideration has been carried out in recent work of Bruch (2014). If the
planet can survive the common envelope phase of the binary, the separation of the planet
from the binary should be larger than the radius RCE of the common envelope. RCE is
approximated by the radius Rmax of progenitor star which evolves to be a giant star with its
core mass equal to the mass of the white dwarf. According to Joss et al. (1987), the relation
between the core mass and the radius is
R ≈
3.7 × 103µ4
1 + µ3 + 1.75µ4 R⊙.
(9)
With the mass of the white dwarf adopted to be 0.6 M⊙, we can estimate the radius of the
giant star in the common envelope phase to be Rmax = 1.53 AU. As the mass loss of the
primary star during the common envelope phase leads to a widening of the planetary orbit,
we should compare the radius of the progenitor star to the original separation of the planet
from the inner binary. According to Zhao et al. (2012), the initial-final-mass-relation for the
white dwarf is
Mf = (0.452 ± 0.045) + (0.073 ± 0.019)Mi.
(10)
Where Mi is the initial mass of the white dwarf, i.e., the mass of the progenitor star, Mf
is the mass of the white dwarf after the common envelope phase. As Mf here is 0.6 M⊙,
the mass of the progenitor is about Mprog = 2.03 M⊙. We assume that the orbital angular
momentum of the planet is constant during the common envelope phase and the mass of the
secondary star changes little. So,
mp
Mprog + M2
Mprog + M2 + mp
q(Mprog + M2)ai = mp
M1 + M2
M1 + M2 + mp
q(M1 + M2)af ,
(11)
where ai is the original semi-major axis of the planet and af is the current semi-major axis
which is determined to be af = 9.12 ± 4.03 AU. At the lower limit, ai ≃ 1.97 AU, at the
upper limit, ai ≃ 5.08 AU, therefore, ai > Rmax even at the lower limit. So the planet can
– 17 –
survive the common envelope phase and have no conflicts with the evolution of the binary
system.
5. Conclusions
The CV LX Ser was observed on four nights using the 1.0 m Cassegrain telescope at
Weihai Observatory of Shandong University, and four new eclipsing times with high precision
were determined. Based on the data from AAVSO International Data base, 58 eclipse times
were redetermined. Further more, combining all the eclipse times from O-C Gateway, we
analyzed the O − C behavior of LX Ser. The O − C diagram shows a cyclic variation with
an amplitude of 0.00035 days and period of 22.8 yr.
For the cyclic variation in the O −C diagram, both of the Applegate mechanism and the
light travel time effect are considered. The energy required to produce the cyclic modulation
of LX Ser was calculated, we found that the minimum value of the required energy is nearly
two times of the total radiated energy of the secondary red dwarf over the whole modulation
period. Therefore, the Applegate mechanism is too feeble to explain the cyclic modulation.
According to the investigation of the light travel time effect, a giant planet with a mass of
7.5 MJ up at a distance of 9.12 AU was explored. Exoplanets can survive almost anywhere
based on the analysis during the last 24 years (e.g., Wolszczan & Frail 1992; Silvotti et al.
2007; Qian et al. 2009, 2012). The existence of the giant planet orbiting the CV LX Ser
is reasonable. By considering all the factors, we analyzed the the stability of the multiple
systems and we found that the outer planet will move slowly away from the centre eclipsing
CV and the system will stay stable. We also considered whether the existence of the giant
planet is compatible with the evolution of LX Ser from a wide orbit binary to a CV and we
determined that the planet can survive the common envelope phase and have no conflicts
with the evolution of the binary system. In the future, more eclipse times with high precision
are needed to confirm the possibility of the giant planet.
This work is partly supported by the National Natural Science Foundation of China
and the Chinese Academy of Sciences joint fund on astronomy (No. U1431105) and by the
National Natural Science Foundation of China (No. 11333002), and by the Natural Science
Foundation of Shandong Province (No. ZR2014AQ019), and by Young Scholars Program of
Shandong University, Weihai (No. 2016WHWLJH07), and by the Open Research Program
of Key Laboratory for the Structure and Evolution of Celestial Objects (No. OP201303).
The authors thanks the AAVSO International Database very much for the observations of
LX Ser. The data used in this paper were contributed by worldwide observers as follows:
– 18 –
Bruno, Alain; Carreno, Alfonso; Cook, Michael; Darriba Martinez, Adolfo; Bubovsky, Pavol;
Foster, James; Gomez, Tomas; Graham, Keith; Gualdoni, Carlo; James, Robert; Macdonald,
Walter; Menzise, Kenneth; Poyner, Gary; Roe, James. Many thanks to the referee for the
helpful comments and suggestions that helped to greatly improve this paper.
REFERENCES
Agerer, Franz, Hubscher, Joachim 2003, IBVS, 5484, 1
Africano, J. L., Klimke, A. 1981, IBVS, 1969, 1
Applegate, J. H. 1992, ApJ, 385, 621
Bonnell, I. A., Bate, M. R. 1994, MNRAS, 269, L45
Burrows, A., Hubbard, W. B., Lunine, J. T., Liebert, J. 2001, RvMP, 73, 719
Brinkworth, C. S., Marsh, T. R., Dhillon, V. S., Knigge, C. 2006, MNRAS, 365, 287
Bruch, Albert 2014, A&A, 566, 101
Chabrier, G., Baraffe, I. 2000, ARA&A, 38, 337
Cox A. N., 2000, Allen's astrophysical quantities, 4th ed. Publisher: New York: AIP Press;
Springer
Dai, Z. B., Qian, S. B. 2009, A&A, 503, 883
Dai, Z. B., Qian, S. B., Fern´andez Laj´us, E. 2009, ApJ, 703, 109
Dai, Z. B., Qian, S. B., Li, L. J. 2013, ApJ, 207, 22
Diethelm, Roger 2003, IBVS, 5438, 1
Eason, E. L. E., Worden, S. P., Klimke, A., Africano, J. L. 1984, PASP, 96, 372
Eastman, Jason, Siverd, Robert, Gaudi, B. Scott 2010, PASP, 122, 935
Hall, D. S. 1989, SSRv, 50, 219
Holman, M. J., Wiegert, P. A. 1999, AJ, 117, 621
Horne, K. 1980, ApJ, 242, L167
Hu, S. M., Han, S. H., Guo, D. F., Du, J. J. 2014, RAA, 14, 719
Hubscher, Joachim 2016, IBVS, 6157, 1
Hubscher, Joachim, Paschke, Anton, Walter, Frank 2005, IBVS, 5657, 1
Hurley, J., Tout, C.-M. 2002, MNRAS, 329, 897
Joss, P. C., Rappaport, S., Lewis, W. 1987, ApJ, 319, 180.
– 19 –
Knigee, C., Baraffe, I. 2011, ApJS, 194, 28
Krajci, Tom 2005, IBVS, 5592,1
Krajci, Tom 2006, IBVS, 5690,1
Liao W.-P., Qian S.-B. 2010, MNRAS, 405, 1930
Lenz, P., Breger, M. 2005, CoAst, 144, 41
Marin, Eduardo, Shafter, A. W., Misselt, K. A. 2007, AAS, 211, 5108
Pringle, J. E. 1975, MNRAS, 170, 633
Potter, Stephen B., Romero-Colmenero, Encarni, Ramsay, Gavin et al. 2011, MNRAS, 416,
2202
Qian, S. B., Han, Z. T., Fern´andez Laj´us, E., Zhu, L. Y., Li, L. J., Liao, W. P., Zhao, E. G.
2015, ApJS, 221, 17
Qian, S.-B., Dai, Z.-B., Liao, W.-P., Zhu, L.-Y., Liu, L., Zhao, E. G. 2009, ApJ, 706, L96
Qian, S.-B., Liao, W.-P., Zhu, L.-Y., Dai, Z.-B., Liu, L., He, J.-J., Zhao, E.-G., Li, L.-J.
2010b, MNRAS, 401, 34
Qian, S.-B., Liu, L., Zhu, L.-Y., Dai, Z.-B., Fernndez Laj´us, E., Baume, G. L. 2012, MNRAS,
422, 24
Qian, S.-B., Liao, W.-P., Zhu, L.-Y., Dai, Z.-B. 2010a, ApJ, 708, L66
Silvotti, R., Schuh, S., Janulis, R. et al. 2007, Nature, 449, 189
Stepanian, J. A. 1979, IBVS, 1630, 1
Wolszczan, A., Frail, D. A. 1992, Nature, 355, 145
Young, P., Schneider, D. P., Shectman, S. A. 1981, ApJ, 244, 259
Zejda, M., Mikulasek, Z., Wolf, M. 2006, IBVS, 5741, 1
Zejda, Miloslav 2004, IBVS, 5583, 1
Zhao, J. K., Oswalt, T. D., Willson, L. A., Wang, Q., Zhao, G. 2012, A&A, 746, 144
This preprint was prepared with the AAS LATEX macros v5.2.
– 20 –
s
s
a
m
y
r
a
m
i
r
p
0.6004
0.6002
0.6
0
0.4
0.2
0
0
0.2
0.1
0
0
s
s
a
m
y
r
a
d
n
o
c
e
s
)
d
(
d
o
i
r
e
p
2
4
6
8
10
12
14
2
4
6
8
10
12
14
2
4
6
8
10
12
14
time(year)
16
x 109
16
x 109
16
x 109
Fig. 7.- Evolution of masses of the two components and period of LX Ser.
|
0909.0505 | 1 | 0909 | 2009-09-02T22:55:23 | On Hydrodynamic Motions in Dead Zones | [
"astro-ph.EP"
] | We investigate fluid motions near the midplane of vertically stratified accretion disks with highly resistive midplanes. In such disks, the magnetorotational instability drives turbulence in thin layers surrounding a resistive, stable dead zone. The turbulent layers in turn drive motions in the dead zone. We examine the properties of these motions using three-dimensional, stratified, local, shearing-box, non-ideal, magnetohydrodynamical simulations. Although the turbulence in the active zones provides a source of vorticity to the midplane, no evidence for coherent vortices is found in our simulations. It appears that this is because of strong vertical oscillations in the dead zone. By analyzing time series of azimuthally-averaged flow quantities, we identify an axisymmetric wave mode particular to models with dead zones. This mode is reduced in amplitude, but not suppressed entirely, by changing the equation of state from isothermal to ideal. These waves are too low-frequency to affect sedimentation of dust to the midplane, but may have significance for the gravitational stability of the resulting midplane dust layers. | astro-ph.EP | astro-ph |
Id: ms.tex 1051 2009-08-27 00:27:01Z joishi
On Hydrodynamic Motions in Dead Zones
Jeffrey S. Oishi
Department of Astronomy, 601 Campbell Hall, University of California at Berkeley,
Berkeley, CA, 94720-3411
[email protected]
and
Mordecai-Mark Mac Low
Department of Astrophysics, American Museum of Natural History, 79th Street at Central
Park West, New York, NY 10024-5192
[email protected]
ABSTRACT
We investigate fluid motions near the midplane of vertically stratified ac-
cretion disks with highly resistive midplanes. In such disks, the magnetorota-
tional instability drives turbulence in thin layers surrounding a resistive, stable
dead zone. The turbulent layers in turn drive motions in the dead zone. We
examine the properties of these motions using three-dimensional, stratified, lo-
cal, shearing-box, non-ideal, magnetohydrodynamical simulations. Although the
turbulence in the active zones provides a source of vorticity to the midplane, no
evidence for coherent vortices is found in our simulations. It appears that this
is because of strong vertical oscillations in the dead zone. By analyzing time
series of azimuthally-averaged flow quantities, we identify an axisymmetric wave
mode particular to models with dead zones. This mode is reduced in amplitude,
but not suppressed entirely, by changing the equation of state from isothermal to
ideal. These waves are too low-frequency to affect sedimentation of dust to the
midplane, but may have significance for the gravitational stability of the resulting
midplane dust layers.
Subject headings: protoplanetary disk,magnetohydrodynamics
-- 2 --
1.
Introduction
A long standing problem in star formation concerns the accretion of high-angular mo-
mentum material in disks around protostars. The rediscovery of the magnetorotational
instability (MRI) in an astrophysical context, along with the discovery of the favorable
transport properties of the turbulence that results from it (Hawley et al. 1995) led to the
widespread belief that a solution to this puzzle had been found. However, the instability
requires a critical coupling strength between field and fluid (Blaes & Balbus 1994; Jin 1996).
Noting the high densities and low temperatures likely at the midplanes of protoplanetary
disks, Gammie (1996) developed a model including a weakly-coupled, MRI-stable "dead
zone" at the midplane of such disks. Since this prediction, considerable effort has gone into
elucidating the conditions under which such a model might exist. Most of this work has
centered on chemical studies of the conditions under which such a dead zone might form
and what its radial and vertical extent might be. Recently, a number of numerical studies,
pioneered by Fleming & Stone (2003), have begun to explore the dynamical consequences
of the dead zone model Fromang & Papaloizou (2006). Turner et al. (2007) have developed
an interesting coupled chemical and magnetohydrodynamical (MHD) model in which the
dead zone disappears because of vertical mixing of ionized gas to the midplane. In their
picture, the dead zone is absent in the sense that large scale magnetic fields with positive
Maxwell stress, (cid:104)−BrBφ(cid:105), form in the midplane and transport significant amounts of angular
momentum even though the midplane is still MRI stable.
Aside from acting as a conduit for material flowing onto the protostar via accretion,
protoplanetary disk midplanes are also the sites of planet formation. One of the outstanding
issues in planet formation is the collection of micron-sized dust into planetesimals, rocky
objects of order 1 km in size. In this context, turbulent fluid motions more general than the
correlated fluctuations that drive accretion become very important. This is because drag
forces couple the dust and gas. Fully MRI-active simulations, Johansen & Klahr (2005)
find preferential concentration of dust in turbulent over-pressures. However, these same
turbulent motions increase the velocity dispersion of the dust grains, possibly leading to de-
structive collisions. The dust-gas coupling can drive strong instability and clumping (Youdin
& Johansen 2007; Johansen & Youdin 2007), which, in combination with MRI turbulence,
allows direct formation of planetesimals by gravitational instability (Johansen et al. 2007).
Together, these results suggest that turbulence plays a complex role in planetesimal forma-
tion. As a first step to considering more complex dust-gas models including dead zones, it
is critical to characterize the fluid motions in the MRI-stable midplane.
Here, we address two specific questions about the fluid motions in dead zones: do
coherent, two-dimensional (2D) vortices form in dead zones? And, what are the wave-like
-- 3 --
motions seen in our previously published models of dead zone dynamics (e.g. Oishi et al.
2007)? Vortices have been studied both for their ability to drive purely hydrodynamic
accretion Umurhan & Regev (2004) and their ability to trap dust particles and thus act as
sites of enhanced planetesimal formation (Bracco et al. 1999; Johansen et al. 2004). Johnson
& Gammie (2005b, hereafter JG05) studied vortex dynamics in a 2D, compressible, shearing
sheet model and concluded that random vorticity perturbations do form coherent large-scale
vortices, but that energy in such structures decays away as t−0.5. They also suggested that
turbulent overshoot from the active layers might provide a vorticity source at the midplane.
Using three-dimensional (3D), compressible but unstratified simulations, Shen et al. (2006)
showed that small amplitude vortices are unstable to elliptical instabilities (Kerswell 2002)
and that finite-amplitude trailing waves are torn apart by a Kelvin-Helmholtz like instability
before they have a chance to form coherent vortices. The anelastic simulations of Barranco &
Marcus (2005) show large, coherent vortices forming at 1 -- 3 H from the midplane, where H
is the disk scale height. This unusual and quite unexpected result suggests that stratification
plays an important role in disk vortex dynamics. Because our time-steps are Courant-limited
by the Alfv´en velocity, we restrict our domain to z < 2H to avoid low-density regions with
extremely high Alfv´en speeds. Nonetheless, we study the effect of stratification on the flow
inside and outside of the dead zone with the goal of understanding if MRI turbulence in
active layers can sustain vortices.
In section 2 we briefly describe our numerical techniques and how we tested them, while
section 3 contains the results of our 3D simulations, and section 4 reviews aspects of vortex
dynamics and presents relevant 2D simulations. Finally, we discuss our results and present
conclusions in section 5.
2. Numerical Method
2.1. Model Description
We use the Pencil Code (Brandenburg & Dobler 2002) to solve the resistive MHD
equations in the vertically stratified, shearing box limit (Hawley et al. 1995).
All models are 3D, isothermal, non-ideal MHD, except for one ideal MHD control run,
and one non-ideal, non-isothermal run (see section 3.3.1). They are run on a domain of
1H × 4H × 4H, where H is the disk scale height. The numerical method is stabilized with
sixth-order hyperviscosity, hyperresistivity, and a hyperdiffusive operator on the density
as described in Oishi et al. (2007). All the models are initialized with a MRI unstable
zero-net flux initial magnetic field given by Bz(x) = B0sin(2πx/Lx) with initial plasma
-- 4 --
β = 2µ0Pg/B2
shearing periodic boundaries in x.
0 = 400. The domain has periodic boundary conditions in y and z, and
The non-ideal models use one of two z-dependent resistivity profiles. One is motivated
by balancing cosmic ray-ionization to a critical depth in column density given by Fleming &
Stone (2003). These models were previously described in Oishi et al. (2007) and are hereafter
referred to as "FS runs". These models were run with cubical zones of size 32 and 64 zones
per scale height. Unless otherwise noted, we use a resolution of 64 zones per scale height
for the FS runs. The second profile, described below, uses hyperbolic tangent functions and
models using it were run with 32 zones per scale height.
2.2. Hyperbolic Tangent Resistivity Profile
This alternate profile offers a sharper transition between active and dead zones than the
FS profile, and allows the thickness of the dead zone to be controlled independently from
its depth. Though simpler, it is less physically motivated than the FS profile. The profile is
given by
(cid:18)
(cid:18) z + z0
(cid:19)
(cid:19)(cid:19)
(cid:18) z − z0
∆z
η(z) =
η0
2
tanh
∆z
− tanh
(1)
where η0 is the midplane resistivity, z0 is the transition height and ∆z is the width of the
transition region, set in this work to ∆z = 5dx, where grid zones have size dx. We set
η0 = 1.67 × 10−5, corresponding to magnetic Reynolds number ReM = c2
s/(ηmidΩ) = 30 at
z < z0. This gives all our tanh models the same value of ReM in the dead zone as our
fiducial FS run has at z = 0.
The Maxwell stresses as a function of height for both sets of profiles are given in Figure 1.
The figure emphasizes the main utility of the tanh profile: because the dead zone resistivity
is independent of its size, we can study larger dead zones at a given resolution.
2.3. Tests
In the numerical study of non-axisymmetric perturbations in disks, a series of test
problems are becoming standard. The vortical and non-vortical shearing waves have closed
analytic solutions (in the former case, a WKB approximation) derived by Johnson & Gammie
(2005a). These waves have been used by JG05 to test a code derived from ZEUS (Stone &
Norman 1992) and the higher-order Godunov code ATHENA (Shen et al. 2006). Aside from
providing an analytic test solution for comparison, the incompressible wave is particularly
-- 5 --
useful for determining the amount of aliasing present in the numerical scheme. As a wave
swings from leading to trailing, it wraps toward axisymmetry. However, in doing so the
number of zones resolving the wave must drop. When the number of zones resolving each
wavelength drops below some threshold, the code may unphysically transfer ("alias") power
from the trailing wave into a new leading wave that will again swing amplify. This danger
is highlighted by both Umurhan & Regev (2004) and Shen et al. (2006), as such power may
spuriously drive coherent vortices and angular momentum transport.
The Pencil Code uses spatially-centered finite differences and Runga-Kutta timestep-
ping, meaning that its discretization lacks any formal algorithmic viscosity. Thus, it should
perform well on this test. Ultimately, however, this property works against the code: without
an explicit viscous term, it does not dissipate energy that moves to smaller scales, where an
unphysical buildup occurs. In this sense, the shearing waves mimic the effect of a turbulent
cascade: as time progresses, kx(t) decreases and the waves become less and less resolved.
This is not a turbulent cascade of course, but the code is nonetheless forced to resolve smaller
and smaller structures, leading eventually to a crash. To avoid this in these tests, as in our
actual runs, we use a sixth-order hyperviscosity scaled by ν6 for stability, setting the hyper-
viscous Reynolds number at the grid scale Re6 = udx5/ν6 (cid:39) 1, where u is the maximum
velocity on the grid.
Figure 2 shows the incompressible shearing wave for runs with resolution N = 64, 128, 256,
demonstrating the ability of the Pencil code to accurately reproduce the analytic solution.
The aliasing time in terms of the orbital frequency is (Shen et al. 2006)
(cid:18) Nx
(cid:19)
taliasΩ0 =
n
q
ny −
kx0
ky
,
(2)
where n is an integer, q = d ln Ω0/d ln r is the shear parameter, and ny = kyLy/2π is the
dimensionless azimuthal wavenumber of the wave. The oscillations around t = 20 appear
to be small amplitude sound waves present in our compressible solution but not in the
incompressible analytic solution.
In Figure 2, the N = 64, 128, and 256 models run to tΩ0 = 100, where Ω0 is the
local orbital frequency. This is well beyond the expected aliasing time taliasΩ0 (cid:39) 45.3n
for Nx = 128 and 90.6n for Nx = 256. The figure makes an interesting point: aliasing
does indeed occur for 642, but because the hyperviscosity is a strongly non-linear function
of resolution, the spurious energy injected by aliasing becomes trivial before the first 1282
aliasing time. Even at 642, each successive aliased wave has less power, suggesting that
sustained vortex activity is not likely to be powered by this numerical effect for long times.
There is no aliasing at all at 1282 and 2562 resolutions, because the hyperviscosity damps the
signal to machine precision before the first aliasing time. We have additionally performed a
-- 6 --
convergence test using resolutions of 64, 128, and 256. Using hyperviscosity, the average error
over tΩ < 10 reaches a constant value at 1282 and does not further converge. However, when
we stabilize the algorithm using a physical Laplacian viscosity instead of hyperviscosity,
fν = ν∇2u, the error converges at roughly ∼ N−1.6 where N is the resolution. We also
consider the compressible shearing wave. Figure 3 demonstrates the code's ability to resolve
the wave with small error to late times with moderate resolution.
Finally, we successfully tested against the 3D, hydrodynamic, non-linear solution of the
shearing box equations derived by Balbus & Hawley (2006). Figure 4 shows the kinetic
energy of the wave as function of time. In this test, as in the others, the wavenumber drops,
so that at larger times the error necessarily grows.
The incompressive shear wave test gives us a strict metric for aliasing: for the lowest
resolution, we used H = 64dx, and saw trivial amounts of aliasing. Doubling the resolution
to H = 128dx, we see no aliasing. Because the amount of vorticity increases in our MHD
simulations when we increase the resolution from H = 32dx to 64dx (see figure 5), we are
confident that any sustained vorticity in our MHD simulations does not come from numerical
aliasing, and that the code accurately tracks any vortical motions present.
3. Results
The excitation and saturation of the MRI occur by t/torb = 5 and the dead zone is
clearly defined by orbit ∼ 10 (see section 3.3 for details). Here, we will restrict our analysis
to times later than 25 orbits in order to avoid transients. Figure 6 shows time series of
vorticity (ω = ∇ × u), and kinetic and magnetic energies. Although the saturated kinetic
energy drops by over an order of magnitude between the ReM = ∞ and the ReM = 3
models, the vorticity only drops by a factor of a few. This can be explained by the presence
of residual vorticity in the dead zone.
3.1. Vorticity and Flow Dimensionality
The MRI produces significant amounts of vorticity, much of which is retained when a
dead zone is introduced. Here, we investigate whether or not this vorticity can coalesce into
coherent vortices that could be significant as a natural environment in which to trap dust
and thus accelerate protoplanet formation.
-- 7 --
In a thin accretion disk, the vertical gravitational acceleration is given by
g = −Ω2zz,
(3)
where Ω is the orbital frequency. In the local approximation, we take Ω0 = Ω(R) where
R is the assumed central radius of the shearing box. For an isothermal disk in hydrostatic
equilibrium, the Brunt-Vaisala frequency of buoyant vertical oscillation is
N 2 = −
g
ρ
∂ρ
∂z
=
0z2
Ω2
H 2 .
(4)
We can diagnose how stratified the flow is by comparing the vertical component of the
vorticity ωz to the Brunt-Vaisala frequency N in a version of the internal Froude number,
(e.g., Barranco & Marcus 2005)
Fr = ωz/(2N ).
(5)
N vanishes at the midplane, and thus Fr formally diverges as z → 0. Nonetheless, Fr provides
a useful diagnostic for the degree of stratification in the flow: when Fr< 1, internal gravity
waves rapidly homogenize vertical disturbances in the vortex, so the system is strongly
stratified, and acts effectively as a 2D flow in which energy cascades to larger scales, similarly
to the Earth's atmosphere. Thus, Fr is an effective measure of the dimensionality of a vortex
in the disk plane.
Previous results show rapid vortex growth in purely 2D systems integrated over z
(Umurhan & Regev 2004; Johnson & Gammie 2005b), and at F r < 1 in 3D stratified
systems (Barranco & Marcus 2005). Figure 7 shows the vertical behavior of F r in our FS
models. All but one of our simulations are purely isothermal, with P = c2
sρ, so gravity waves
themselves are excluded. However, because the basic dynamical properties of the MRI are
not very sensitive to the equation of state (Stone et al. 1996), we do not expect that including
gravity waves will change the F r profile. We confirm this expectation in figure 7, which also
includes our control ReM = 30 run with a non-isothermal, γ = 5/3 ideal gas equation of
state (see section 3.3.1 for more details). The F r profile is roughly similar to the run with
isothermal ReM = 30.
In our ideal MHD run the flow is effectively 3D through almost the entire domain -- the
MRI is a strong enough vorticity source to overwhelm the homogenizing effect of vertical
oscillations almost everywhere. On the other hand, Figure 7 suggests that the dead zone
models have an effectively 2D flow at nearly all heights, most particularly in the dead zone
of the ReM = 3 model. Near the midplane, F r diverges, as the stratification there goes to
zero, so we would expect the very near midplane region to act as a 3D flow, consistent with
the Barranco & Marcus (2005) results showing disappearance of their imposed vortices in
that region.
-- 8 --
However, Figure 8 shows that no vortices form in either the dead or the active zone. In
the dead zone, the vorticity remains confined to elongated, nearly axisymmetric bands. In
fact, no difference is seen in ωz between heights within the dead zone -- only between dead and
active zones. Thus, our results combined with Shen et al. (2006) and Barranco & Marcus
(2005), show that even in the presence of a vorticity source such as active zones, stratification
is necessary but insufficient to produce vortices.
Table 1 shows the distribution of enstrophy among x,y, and z components for FS models
and the ideal MRI run. A strongly two-dimensional flow shows a preferred direction (Bran-
denburg et al. 1995). The ideal MRI run and the active zones of the FS models agree well
with an isotropic distribution of vorticity. The dead zones show a strong preference for the y
direction, indicating a circulation in the x − z plane, as indeed images bear out (see below).
The coherent anticyclonic vortices we set out to investigate would have shown a strong z
component.
Figure 9 shows the components of velocity dispersion for each of the FS runs as a
function of time, computed separately inside and outside of the dead zone. The z component
of velocity dominates within the dead zone, while outside of it, it is a factor of a few smaller
than the others. (Note that the oscillation in the dead zone velocities is real, but the plot
was made with relatively crude time resolution, δt = 0.5torb and some aliasing from higher
frequency signals is certainly present.) The dead zone motion is clearly not dominated by
horizontal flows.
3.2. Morphology and velocity vectors
Figure 10 shows images of density overlaid with velocity vectors for the midplane (the
In
x-y plane at z = 0) and Figure 11 the x-z plane at y = 0 for each of the FS runs.
the dead zone midplanes (the right three panels), compressive waves dominate the velocity
field: velocity vectors lie almost entirely orthogonal to the density striations in the x-y plane.
However, in the x-z plane, inertial mode signatures appear: the velocity is dominated by the
z component, traveling up and down in well defined vertical bands.
The dead zone shows ordered uz alternating between positive and negative values across
the dead zone that become more pronounced as ReM decreases (Fig. 11, right three panels).
This suggests that the vertical oscillations are at θ = 0, with energy transport from gravity
waves purely vertical between the active zones and the midplane. This is in stark contrast to
the fully MRI-active case. Near its midplane, there are no well-defined vertical oscillations,
and the ReM = 100 model does not show nearly as well-ordered motion as the larger dead
-- 9 --
zone models. This is perhaps not surprising given that the fastest growing MRI mode has a
growth rate qM RI ∼ Ω0, which is greater than N until around a scale height.
3.3. Wave Modes
The MRI is not operating in the densest parts of the disk in our dead zone models, so
we expect that the motions excited in this region will take the form of linear wave modes
stochastically excited by the active zone turbulence. The following analysis is motivated
by the low-frequency oscillation in the volume-averaged kinetic energy most clearly shown
in the ReM = 3 data between 60 < t < 100 in Figure 6. Fourier analysis of this kinetic
energy time series for runs with a dead zone show a clear peak in frequency space that shifts
slightly to lower frequency with increasing dead zone size, suggesting the presence of a large
amplitude coherent oscillation in these runs.
In order to better understand these dead zone oscillations, we take temporal power spec-
tra of the radial and vertical velocity components ux, and uz, and the density perturbation
δρ = ρ(x, y, z) − ρ0(z). Arras et al. (2006) and Brandenburg (2005) used similar methods to
diagnose global modes of oscillation in fully magnetorotationally turbulent disks. The for-
mer were able to isolate acoustic and inertial oscillations. However, the latter found no clear
wave signatures for MRI turbulence, though they did recover acoustic and inertial modes for
forced hydrodynamic turbulence in a shearing box.
Figure 12 shows spectra computed by taking an FFT of a time series of each variable
averaged in the y-direction at every (x, z) point on the grid of our low resolution model. The
resulting power spectra were then averaged to raise signal-to-noise. Azimuthal averaging
allows us to eliminate much of the MRI power in the active zones (Arras et al. 2006), though
this may not be a significant source in some of our larger dead zone models. The dominant
peak in all variables at max (cid:39) 0.23Ω0 shifts only very slightly, to lower frequency, as the
dead zone is made thicker by a factor of two in the tanh runs. This suggests that the vertical
thickness of the dead zone is not setting the oscillation frequency. We have also determined
that this characteristic frequency is unaffected by the radial (x) extent of the box by running
the ReM = 30 model with Lx = 2 and confirming that the oscillation frequency is unchanged.
Because the box is periodic, standing oscillations can be excited, with discrete frequen-
cies. The lowest frequency acoustic mode is in the vertical direction and has a frequency
(cid:39) cs2π/Lz = 1.11Ω0, which is clearly larger than max. The isothermal equation of state
used in these models precludes gravity waves (which would also have < Ω0), and so we
tentatively identify them as an inertial oscillation. Furthermore, the small shift seen in the
-- 10 --
tanh runs as the dead zone thickness increases is actually to lower frequency, while a higher
boundary would drive internal oscillations at a higher frequency (see Eq. 4) .
3.3.1. Non-isothermal models and Buoyancy Forces
The low frequency of these oscillations is suggestive of a gravity mode. However, these
sρ with cs constant everywhere, precluding buoyant re-
runs are strictly isothermal, P = c2
sponses and thus gravity waves.
Recently, Bai & Goodman (2009) noted that isothermal equations of state might spu-
riously enhance vertical mixing, as buoyant forces can inhibit vertical mixing. Likewise,
the lack of buoyancy in our isothermal models may artificially enhance both the residual
transport at the midplane and the coherent oscillations. In order to understand the effect of
isothermality on these issues, we ran a single model at 32 zones per scale height including
an ideal gas equation of state P = ρkT /µ and an entropy equation,
ρT (
∂s
∂y
y
∂s
∂t
+ u · ∇s + u0
) = η(z)µ0j2 + ζρ (∇ · u)2 + κ6∇6s,
where κ6 is a hyperdiffusivity of entropy, u0
y = −3/2Ωx is the Keplerian velocity profile, Ω is
the rotation rate of the shearing box, and all other symbols have their usual meanings. The
entropy equation includes resistive and shock heating but not heating from the hyperdiffu-
sivities (these terms are small, and do not significantly contribute to the thermal balance).
We chose a ratio of specific heats γ = 5/3.
(6)
The boundary conditions on this run as on the others are periodic in all directions, which
precludes the escape of heat. We choose this rather unrealistic setup to directly compare
with our isothermal simulations, in order to demonstrate that our results are not due to the
lack of buoyancy in the isothermal case. Because of the turbulent heating, the total energy
in the box increases compared to the isothermal case. Although the turbulence is strongest
in the surface layers, there is no temperature inversion in the vertical direction: the midplane
remains hotter than the surface. The disk quasi-statically adjusts to new equilibrium density
and pressure profiles as the energy increases. While these profiles are somewhat different
from the standard Gaussian density profile expected for accretion disks with linear z-gravity
profiles, Figure 5 shows that the main result of this difference is a factor ∼ 2 increase in
magnetic energy, and a factor ∼ 1.8 increase in kinetic energy.
The overall morphology of the flow is roughly similar to the isothermal model, with a
reduced amplitude of motion in the dead zone but a similar large-scale structure. The stress
profile across the dead zone is narrower but deeper in the ideal gas run than the isothermal
-- 11 --
one because of the pressure confinement caused by turbulent heating in a periodic box absent
cooling. We directly compare the Maxwell and Reynolds stresses for isothermal (labeled
γ = 1) and ideal gas (γ = 5/3) in Figure 13. We find that inclusion of the buoyant forces
reduces the amplitude of the vertical motions and increases their characteristic frequency
to max (cid:39) 0.5Ω0, still well below the first acoustic mode, as shown in Figure 14. However,
the motions are not suppressed, even in this model in which turbulent heating is not at all
balanced by cooling. A real disk in which cooling is allowed will fall between the limits of
the isothermal model and this heating only model. Therefore, we argue that the dead zone
motions represent a physical phenomenon even in the presence of buoyancy forces.
4. Midplane Vortex Dynamics
We see no vortices in our simulations. We must establish that this is a physical result,
and not due to insufficient resolution. JG05 demonstrate that sufficient numerical resolution
is necessary to sustain kinetic energy and angular momentum transport from vortices. In
their 2D simulations, they find 128dx/H to be the critical value, raising the question of
whether vortices can form in 2D at our resolutions (32dx/H or 64dx/H)? Furthermore, the
typical radial size scale of a vortex in their simulations is roughly H, the entire width of our
box. Can vortices form in such cramped quarters?
4.1. Two-Dimensional Vortex Dynamics
In order to clarify these issues, we ran four sets of 2D hydrodynamic simulations. First,
we studied resolution. A comparison between Figure 2 and Figure 1 of JG05 shows that the
sixth-order Pencil Code is significantly less diffusive than their second-order, Zeus-derived
method, suggesting that we might be able to see sustained vortex activity at lower resolution.
Therefore we ran a set of models using their domain with size Lx = Ly = 4H at resolutions
of (16, 32, 64)dx/H, all with Gaussian random velocity perturbations with the same initial
perturbation amplitude, σ = 0.8cs. Figure 15 shows the kinetic energy and α-parameter, α =
(cid:104)Σuxuy(cid:105) for 2D. This Figure demonstrates clear convergence in these integrated quantities
at 32dx/H.
Next, we turn to the question of domain size. JG05 use a domain four times larger in
radial extent than we do in our 3D simulations. Could this inhibit vortex formation in our
MHD runs? To test this, we ran a set of 2D hydrodynamic models with the same domain
as the midplane of our 3D simulations, Lx = 1H and Ly = 4H, again with resolutions of
-- 12 --
(16, 32, 64)dx/H, seeded with perturbations of magnitude σ = 0.8cs. Figure 16 shows that
vortex morphology is largely similar in both domains at three times, though there are more
vortices present in the square domain.
Figure 17 shows the resulting energy and α. Although the results are not as well
converged as in the square domain, there is still evidence for vortex activity with lifetimes
clearly greater than the MRI growth time τMRI ∼ Ω at a resolution of 32dx/H.
What, then, is the reason we do not see vortex activity in the 3D dead zone simulations?
We have argued that dimensionality is not the limiting factor, as regions with lower Froude
number that are more effectively 2D do not show more vorticity. It also appears clear that
the 2D vortex activity found in previous simulations can be reproduced on domains like ours
at the resolutions we use in our MHD runs, so resolution and grid size are also not limiting.
Rather, the discriminant appears simply to be the strength of perturbations required to
trigger long-lived vortices. We ran a third set of 2D hydrodynamic simulations, this time
at a fixed resolution of 32dx//H, but with decreasing velocity perturbation magnitudes
σ = 0.8cs, 0.5cs, and 0.1cs. Figure 18 shows that below σ = 0.8, the kinetic energy and
α drop precipitously after only a few orbits. (Note that we keep the domain size fixed at
Lx = 1H, Ly = 4H, unlike in the tests of JG05, in which the domain size was scaled linearly
with the strength of the velocity dispersion.) We conclude that compressible vortices require
strong σ = 0.8cs initial perturbations in order to survive for many orbits.
Umurhan & Regev (2004) see perturbation energy sustained essentially indefinitely for
Re = ∞ and decaying away only very slowly for Re = 50000, but their model is incompress-
ible. They show this approximation to be rigorously valid for the small length scales they
consider. However, interestingly, they find that the perturbation energy saturates at roughly
(cid:46) 10−2Eshear, where Eshear is the energy in the shear. They define the turbulent intensity of
an incompressible shear flow as
(cid:46) 10
−2,
(7)
(cid:82) u2 dx dy
(cid:82) u2
sh dx dy
ε0 =
with u the disturbance velocity and ush the shear velocity. We write ush = qΩx, integrate
for the Keplerian case q = −3/2, and use our definition of (thermal) scale height H = 2cs/Ω
to arrive at
(cid:18) Lx
(cid:19)2
H
(cid:104)Ma2(cid:105) =
σ2
c2
s
=
3
8
ε0
,
(8)
with (cid:104)Ma2(cid:105)0.5 = Marms the RMS Mach number. Using their saturation value, this gives
Marms ∼ 0.06(Lx/H), or about 0.06 over a domain of H. This is an order of magnitude
below the levels necessary to incite vortices on the larger scales that we simulate in which
compression becomes important. If we assume that small scale, non-linear, transient growth
-- 13 --
instability scales with Reynolds number as found by Lesur & Longaretti (2005), not only is
the incompressible, small-scale, transient growth irrelevant for angular momentum transport,
its saturation level is so small that larger scale vortices in the compressible regime will not
form at all.
4.2. Long-lived axisymmetric structures
Recently, Johansen et al. (2009) reported the existence of long-lived axisymmetric zonal
flows in MRI turbulence, provided the radial width of the box is larger than ∼ 1H. These
structures result from an inverse cascade of magnetic energy to the largest scales present in
the box. This energy drives a large scale Maxwell stress, which in concert with the Coriolis
force enforces the zonal flow. The end result are long-lived (∼ 10s of orbits) pressure and
density fluctuations. It is easy to imagine that these complex structures could have significant
influence on motions in the dead zones, assuming there is sufficient magnetic energy to launch
the zonal flows in the first place. Figure 1 implies that this criterion is unlikely to be met:
even in the moderate sized dead zone models (e.g., ReM = 30 and z0 = 1), the total Maxwell
stress is many orders of magnitude lower than that in the MRI active zones. However, given
that Johansen et al. (2009) demonstrate that the zonal flows are subtle, stochastically driven,
and non-linear, we find such a crude estimate to be unsatisfactory.
Therefore, we re-ran our ReM = 30 FS model with a radially-enlarged domain, running
from −H < x < H. By averaging ρ over the entire y − z domain, including both active
and dead zones, we recover the reported long-lived axisymmetric structures. Nevertheless,
our results remain robust: there is no vortex formation, and figure 19 demonstrates that the
midplane oscillations are quite similar between the larger domain and our fiducial case.
5. Discussion
We find a complete lack of coherent anticyclonic vortices in all of our stratified local
simulations of dead zones, despite the presence of a vorticity source in the form of active
zones. While our experiments do not tell us exactly what happens, it appears that even if
the midplane were a pure 2D layer, large scale vortex formation would not occur because
the amplitude of velocity perturbations driven in the dead zone by the active layers is well
below that needed. Recent work by Lesur & Papaloizou (2009) suggests that vortices in 3D
disks are always parametrically unstable, though the growth rates of such instabilities can
be very slow. Even in the presence of an active vorticity source, though, we find no vortex
-- 14 --
formation.
Our simulations cover a smaller range of z than Barranco & Marcus (2005), and it is
possible that the increased stratification present at higher z may change our results. To
strengthen the results presented here, direct comparison with Barranco & Marcus (2005)
including compressibility would be quite enlightening. Furthermore, almost all of our simu-
lations use an isothermal relation for pressure and density. This forces density gradients and
pressure gradients to coincide, eliminating the baroclinic term driving the vortex formation
from breaking gravity waves in Barranco & Marcus (2005). Relaxing this requirement in
our simulations may aid vortex formation at higher scale heights though our one adiabatic
model without cooling did not show vortex formation either.
However, we do find a large-scale, axisymmetric oscillation about the midplane in all
dead zone models. This oscillation carries nearly all the kinetic energy of the dead zone. We
are unable to associate it with a normal mode using a simple linear dispersion relation. A
more detailed analysis including the effects of continuous stratification could shed consid-
erable light on the situation. Regardless of such details, the low frequency (less than Ω0)
suggests that the mode will not have much effect on dust sedimentation, which is typically
dominated by random fluctuation with correlation times τcorr (cid:39) 0.15torb, corresponding to
frequencies (cid:39) 7Ω0 (Fromang & Papaloizou 2006). It could, in principle, cause a warp in the
dust layer, which in turn could affect the gravitational stability (Goldreich & Ward 1973;
Johansen et al. 2007). Such phenomena can only be examined with a combined treatment
of dust, self-gravity, and a dead zone, which we defer to future work.
Finally, global dynamics may significantly alter vortex formation properties. A real
protoplanetary disk is thin and quasi-2D for horizontal scales greater than the scale height.
It is quite possible that robust, anticyclonic vortices may form within a disk, but our results
suggest that a local mechanism is not responsible. The baroclinic instability (Klahr &
Bodenheimer 2003; Petersen et al. 2007a,b) remains a candidate for forming such large scale
vortices.
JSO would like to thank the Hausdorff Research Institute for Mathematics at Universtat
Bonn, where a considerable portion of this work was completed, for their hospitality and
financial support. J. Hawley, K. Menou, and P. Arras made useful suggestions on an early
version of this manuscript. Computations were performed at the Parallel Computing Facility
of AMNH and the Cray XT3 "Big Ben" at the Pittsburgh Supercomputing Center, the latter
of which is supported by the National Science Foundation. M-MML was partially supported
by NASA Origins of Solar Systems grant NNX07AI74G and by NSF grant AST08-35734.
-- 15 --
REFERENCES
Arras, P., Blaes, O., & Turner, N. J. 2006, ApJ, 645, L65
Bai, X.-N., & Goodman, J. 2009, arXiv:0904.1240
Balbus, S. A., & Hawley, J. F. 2006, ApJ, 652, 1020
Barranco, J. A., & Marcus, P. S. 2005, ApJ, 623, 1157
Blaes, O. M., & Balbus, S. A. 1994, ApJ, 421, 163
Bracco, A., Chavanis, P. H., Provenzale, A., & Spiegel, E. A. 1999, Physics of Fluids, 11,
2280
Brandenburg, A. 2005, Astronomische Nachrichten, 326, 787
Brandenburg, A., & Dobler, W. 2002, Comput. Phys. Comm., 147, 471
Brandenburg, A., Nordlund, A., Stein, R. F., & Torkelsson, U. 1995, ApJ, 446, 741
Gammie, C. F. 1996, ApJ, 457, 355
Fleming, T., & Stone, J. M. 2003, ApJ, 585, 908
Fromang, S., & Papaloizou, J. 2006, A&A, 452, 751
Goldreich, P., & Ward, W. R. 1973, ApJ, 183, 1051
Hawley, J. F., Gammie, C. F., & Balbus, S. A. 1995, ApJ, 440, 742
Jin, L. 1996, ApJ, 457, 798
Johansen, A., Andersen, A. C., & Brandenburg, A. 2004, A&A, 417, 361
Johansen, A., & Klahr, H. 2005, ApJ, 634, 1353
Johansen, A., Oishi, J. S., Mac Low, M-M., Klahr, H., Henning, Th., & Youdin, A. 2007
Nature, 448, 1022
Johansen, A., & Youdin, A. 2007, ApJ, 662, 627
Johansen, A., Youdin, A., & Klahr, H. 2009, ApJ, 697, 1269
Johnson, B. M., & Gammie, C. F. 2005a, ApJ, 626, 978
Johnson, B. M., & Gammie, C. F. 2005b, ApJ, 635, 149 (JG05)
-- 16 --
Kerswell, R. R. 2002, Annual Review of Fluid Mechanics, 34, 83
Klahr, H. H., & Bodenheimer, P. 2003, ApJ, 582, 869
Lesur, G., & Longaretti, P.-Y. 2005, A&A, 444, 25
Lesur, G., & Papaloizou, J. C. B. 2009, A&A, 498, 1
Oishi, J. S., Mac Low, M.-M., & Menou, K. 2007, ApJ, 670, 805
Petersen, M. R., Julien, K., & Stewart, G. R. 2007, ApJ, 658, 1236
Petersen, M. R., Stewart, G. R., & Julien, K. 2007, ApJ, 658, 1252
Shen, Y., Stone, J. M., & Gardiner, T. A. 2006, ApJ, 653, 513
Stone, J. M., & Norman, M. L. 1992, ApJS, 80, 753
Stone, J. M., Hawley, J. F., Gammie, C. F., & Balbus, S. A., ApJ, 463, 656
Turner, J. S. 1973, Buoyancy Effects in Fluids. Cambridge, UK: Cambridge University Press
Turner, N. J., Sano, T., & Dziourkevitch, N. 2007, ApJ, 659, 729
Umurhan, O. M., & Regev, O. 2004, A&A, 427, 855
Youdin, A., & Johansen, A. 2007, ApJ, 662, 613
This preprint was prepared with the AAS LATEX macros v5.2.
-- 17 --
Fig. 1. -- Maxwell stress as a function of height z for three different values of magnetic
Reynolds number using the FS profile, and three different transition heights z0 using the
tanh resistivity profile (Eq. 1). At very high η values, the stress is occasionally positive and
therefore absent on this plot.
"!$#&%'!$#&%)( "!+*
,.- /!+* #0#,.- /!$10#,.-/!$1 -- 18 --
Fig. 2. -- Time evolution of the velocity perturbation amplitude δux of an incompressible,
shearing wave (upper panel) using physical viscosity for resolutions from 642 to 1282. The
analytic solution is given in black. Aliasing is present in the 642 solutions but injects only
trivial amounts of energy. The lower panel shows absolute value of the error, = δux −
δuanalytic
. Note that this error smoothly asymptotes to the analytic solution as the code
hyperviscously damps the wave faster than aliasing can inject spurious energy.
x
020406080100t 10-910-810-710-610-510-1310-1210-1110-1010-910-810-710-610-510-410-3ux64212822562Analytic -- 19 --
Fig. 3. -- Time evolution of the amplitude δuy of a compressible, shearing wave using
hyperviscosity for resolutions from 322 to 1282. The analytic solution is given in black. The
lower panel shows the absolute value of the error, = δuy − δuanalytic
, between the Pencil
Code solution and a numerically integrated solution to the exact parabolic cylinder equation
describing the wave.
y
0246810t 10-1210-1110-1010-910-810-710-62.01.51.00.50.00.51.0uy1e464212822562Analytic -- 20 --
Fig. 4. -- Time evolution of the kinetic energy of a 3D, non-linear, incompressible, shearing
wave using hyperviscosity for resolutions from 643 to 2563. A numerical integration of the
exact wave differential equation is given in black and labeled "Analytic". The lower panel
shows absolute value of error, = Ekin − Eanalytic
kin
0510152025t 10-610-510-410-310-210-310-210-1Ekin/E025631283643Analytic -- 21 --
Fig. 5. -- From top to bottom: volume averaged vorticity (ω), kinetic, and magnetic energy
comparing isothermal and ideal gas ReM = 30 runs.
Isothermal runs at H = 64dx and
H = 32dx are overplotted. The amount of vorticity increases with increasing resolution.
"!#$ % '&()*+-,*+/.
10*+-,3241576 -- 22 --
Fig. 6. -- Same as figure 5 for each H = 64dx resolution FS run.
020406080100t/torb10-410-310-2B2/P010-410-310-2Ekin/P010-1100 /ReM=ReM=100ReM=30ReM=3 -- 23 --
Fig. 7. -- Internal Froude number versus height above the midplane for the FS runs. At
Fr(cid:46) 1, the flow becomes strongly stratified and effectively two-dimensional for vortices lying
in the disk plane.
2.0 1.5 1.0 0.50.00.51.01.52.0z/H0.00.51.01.52.0FrReM=ReM=100ReM=30ReM=3ReM=30,=5/3 -- 24 --
Fig. 8. -- Vorticity in horizontal (x-y) planes as a function of height above and below the
midplane for the ReM = 3 run. The effectively 2D nature of the flow does not appear to
create regions in which coherent vortices can form.
0.50.0.5x/H1.51.00.50.00.51.01.5y/Hz=1.5z=0.5z=0.0z=0.5z=1.5 -- 25 --
In all cases with a
Fig. 9. -- Mean square velocities in each direction for the FS runs.
dead zone, the upper three curves show the active zone and the lower three the dead zone
velocities. Note the transition in dominant component from y velocity to z velocity from
active to dead in all cases.
10-510-410-310-210-1 u2x,y,z/c2sReM=u2xu2yu2zReM=10030405060708090t/torb10-510-410-310-2 u2x,y,z/c2sReM=3030405060708090100t/torbReM=3 -- 26 --
Fig. 10. -- Slices through the midplane of density (color) with arrows following the in-plane
(x-y) velocity. Each density image is scaled to its own minimum and maximum to emphasize
morphology.
−1.5−1.0−0.50.00.51.01.5y/H−0.50.0.5x/HReM=∞ReM=100ReM=30ReM=30.80.9ρ0.951.00ρ0.961.00ρ0.9750.990ρ -- 27 --
Fig. 11. -- Slices through the y = 0 (radial-vertical) plane of density (color) with arrows
following the in-plane (x-z) velocity. Each density image is scaled to its own minimum and
maximum to emphasize morphology. Note the vertical circulation in dead zones viewed in
this plane.
−1.5−1.0−0.50.00.51.01.5z/H−0.50.0.5x/HReM=∞ReM=100ReM=30ReM=30.30.6ρ0.250.50ρ0.250.50ρ0.250.50ρ -- 28 --
Fig. 12. -- Temporal power spectra for ux, uy, ρ. Acoustic modes are present in all runs at
frequencies > 1Ω. Dead zone models also show a low-frequency inertial oscillation apparently
driven by the active zones.
10-910-810-710-6 210-810-710-610-5ux2/c2s10-810-710-610-5uz2/c2sReM=10-1010-910-810-710-610-5 210-910-810-710-610-5ux2/c2s10-910-810-710-610-510-4uz2/c2sReM=3010-1100101/10-1110-1010-910-810-7 210-910-810-710-610-5ux2/c2s10-1010-910-810-710-610-5uz2/c2sz0=110-1100101/10-1010-910-810-710-610-5 210-910-810-710-610-5ux2/c2s10-910-810-710-610-510-4uz2/c2sz0=0.5 -- 29 --
Fig. 13. -- Stress vs z for isothermal and ideal gas runs.
! "#$&%(')%+*-,/.1032#$%'%*-,.10546798;:':*=<>.103278;:':*<.10546 -- 30 --
Fig. 14. -- Temporal power spectra of ux, uy, ρ for ReM = 30 with ideal gas equation of
state (labeled γ = 5/3) and isothermal (γ = 1) runs overplotted. Both inertial and acoustic
waves are significantly damped but still present in the γ = 5/3 case. The large amplitude
mode at Ω (cid:39) 0.2 in the isothermal case remains present, though at reduced amplitude and
increased wavenumber Ω (cid:39) 0.5. The normalization difference between the two spectra is
due to the extra heat retained in the non-isothermal case.
10-1100101 /10-910-810-7210-810-710-610-5ux2/c2s10-810-710-610-5uz2/c2s=5/3=1 -- 31 --
Fig. 15. -- Kinetic energy and α for a 2D domain with Lx = Ly = 4H at three different
resolutions, (16, 32, 64)dx/H, using initial velocity perturbations of magnitude σ = 0.8cs.
Energy and α are clearly well converged at a resolution of 32dx/H.
!"# !$% !&'()*+,+ +
-
.
0/
21
3 -- 32 --
Fig. 16. -- Potential vorticity of the entire x-y plane at times tΩ = 15.7, 31.4, 47.1 (left to
right) upper panel A 4H square domain. Blue tones indicate negative potential vorticity, red
lower
positive. Vortex formation is similar to that found in Johnson & Gammie (2005b).
panel A domain 1H × 4H. In both panels, colors are scaled to the minimum and maximum
of each image to highlight vortex morphology.
-- 33 --
Fig. 17. -- Kinetic energy and α for a domain with Lx = 1H, and Ly = 4H, at three different
resolutions, (16, 32, 64)dx/H, using initial velocity perturbations of magnitude σ = 0.8cs.
Energy and α are not as well converged as in Figure 15, but sustained vortex activity does
seem to occur even at 32 zones/H.
!"# !$% !&'()*+,+ +
-
.
0/
21
3 -- 34 --
Fig. 18. -- Kinetic energy and α for a domain with Lx = 1H and Ly = 4H at resolution
of 32dx/H, using three different initial velocity dispersions, σ = (0.1, 0.5, 0.8)cs. Sustained
vortex activity clearly requires strong initial perturbations.
! #"%$'&! #"%$)(! #"%$*+,-./010 0
2
3
54
76
8 -- 35 --
Fig. 19. -- Temporal power spectra of ux, uy, ρ for ReM = 30 with the standard box width
(−0.5H < x < 0.5H) and a box twice as wide (−H < x < H). The wider box features the
large axisymmetric pressure bumps reported by Johansen et al. (2009), but this has little
effect on the oscillations reported in this paper.
10-1100101 /10-910-810-7210-810-710-610-5ux2/c2s10-810-710-610-5uz2/c2sLx=2HLx=H -- 36 --
Table 1. Directional normalized enstrophy for dead and active zones.
Run
x/ω2
ω2
Active
y/ω2
ω2
z /ω2
ω2
x/ω2
ω2
Dead
y/ω2
ω2
z /ω2
ω2
64Rinf
64R100
64R30
64R3
32z0.5
32z1.0
0.31
0.32
0.35
0.31
0.26
0.27
0.34
0.34
0.36
0.41
0.48
0.46
0.35
0.35
0.29
0.28
0.27
0.26
···
0.11
0.07
0.06
0.10
0.07
···
0.69
0.79
0.68
0.68
0.76
···
0.20
0.14
0.26
0.22
0.17
|
1607.00661 | 2 | 1607 | 2016-11-18T19:34:44 | IAU Meteor Data Center | the shower database: a status report | [
"astro-ph.EP"
] | Currently, the meteor shower part of Meteor Data Center database includes: 112 established showers, 563 in the working list, among them 36 have the pro tempore status. The list of shower complexes contains 25 groups, 3 have established status and 1 has the pro tempore status. In the past three years, new meteor showers submitted to the MDC database were detected amongst the meteors observed by CAMS stations (Cameras for Allsky Meteor Surveillance), those included in the EDMOND (European viDeo MeteOr Network Database), those collected by the Japanese SonotaCo Network, recorded in the IMO (International Meteor Organization) database, observed by the Croatian Meteor Network and on the Southern Hemisphere by the SAAMER radar. At the XXIXth General Assembly of the IAU in Honolulu, Hawaii in 2015, the names of 18 showers were officially accepted and moved to the list of established ones. Also, one shower already officially named (3/SIA the Southern iota Aquariids) was moved back to the working list of meteor showers. At the XXIXth GA IAU the basic shower nomenclature rule was modified, the new formulation predicates \The general rule is that a meteor shower (and a meteoroid stream) should be named after the constellation that contains the nearest star to the radiant point, using the possessive Latin form". Over the last three years the MDC database was supplemented with the earlier published original data on meteor showers, which permitted verification of the correctness of the MDC data and extension of bibliographic information. Slowly but surely new database software options are implemented, and software bugs are corrected. | astro-ph.EP | astro-ph | IAU Meteor Data Center -- the shower database: a status report
Tadeusz Jan Jopeka, Zuzana Kanuchov´ab,∗
aInstitute Astronomical Observatory, Faculty of Physics, A.M. University, Pozna´n, Poland
bAstronomical Institute of the Slovak Academy of Sciences, 05960 Tatransk´a Lomnica, Slovakia
6
1
0
2
v
o
N
8
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
6
6
0
0
.
7
0
6
1
:
v
i
X
r
a
Abstract
Currently, the meteor shower part of Meteor Data Center database includes: 112 established showers, 563 in the working
list, among them 36 have the pro tempore status. The list of shower complexes contains 25 groups, 3 have established
status and 1 has the pro tempore status.
In the past three years, new meteor showers submitted to the MDC database were detected amongst the meteors
observed by CAMS stations (Cameras for Allsky Meteor Surveillance), those included in the EDMOND (European viDeo
MeteOr Network Database), those collected by the Japanese SonotaCo Network, recorded in the IMO (International
Meteor Organization) database, observed by the Croatian Meteor Network and on the Southern Hemisphere by the
SAAMER radar.
At the XXIXth General Assembly of the IAU in Honolulu, Hawaii in 2015, the names of 18 showers were officially
accepted and moved to the list of established ones. Also, one shower already officially named (3/SIA the Southern iota
Aquariids) was moved back to the working list of meteor showers. At the XXIXth GA IAU the basic shower nomenclature
rule was modified, the new formulation predicates "The general rule is that a meteor shower (and a meteoroid stream)
should be named after the constellation that contains the nearest star to the radiant point, using the possessive Latin
form".
Over the last three years the MDC database was supplemented with the earlier published original data on meteor
showers, which permitted verification of the correctness of the MDC data and extension of bibliographic information.
Slowly but surely new database software options are implemented, and software bugs are corrected.
Keywords: Meteoroid streams; Meteor showers; Established meteor showers; IAU MDC; Meteor database; Meteor
showers nomenclature rules.
1. Introduction
The Meteor Data Center shower database (MDC) came
into being in 2007 (Jenniskens, 2008; Jopek and Jen-
niskens, 2011; Jopek and Kanuchov´a, 2014) as a comple-
ment to the already existing orbital part of the IAU Meteor
Data Center (Lindblad, 1987, 1991; Lindblad and Steel,
1994; Lindblad et al., 2003; Svoren et al., 2008; Porubcan
et al., 2011; Neslusan et al., 2014).
The main objective of the MDC database is to archive
information on meteor showers: their geocentric and he-
liocentric parameters. Before inclusion into the MDC, a
new meteor shower obtains a unique name and code, ac-
cording to a set of officially formalized rules. One of the
results of this procedure is a considerably less confusion in
the meteor shower names published in literature.
The MDC showers database is posted on the Web on
two servers: at the Astronomical Observatory, A. M. Uni-
versity in Pozna´n, Poland (http://localhost/~jopek/
MDC2007/) and at the Astronomical Institute of the Slovak
∗Corresponding author
Email address: [email protected] (Zuzana Kanuchov´a)
Preprint submitted to Journal of LATEX Templates
Academy of Sciences, 05960 Tatransk´a Lomnica, Slovakia
(https://www.ta3.sk/IAUC22DB/MDC2007/).
Beyond the efficient collection, checking and dissemina-
tion of the meteor showers data, the main task of the MDC
(in the conjunction with the Working Group on Meteor
Shower Nomenclature of IAU Commission F1) is to for-
mulate a descriptive list of the established meteor showers
that can receive official names during the upcoming IAU
General Assembly.
The first Working Group (named at that time the Task
Group) was appointed at the XXVIth GA IAU in 2006 held
in Prague. The activity of each WG lasts three years in the
period between two IAU General Assemblies. At present,
starting from the XXIXth GA IAU in Honolulu, the cur-
rent members of the WG in the 2015-2018 triennium are:
Diego Janches (chair), Peter Brown, Peter Jenniskens,
Tadeusz J. Jopek (editor of the MDC shower database),
Zuzana Kanuchov´a, Gulchekhra I. Kokhirova, Masahiro
Koseki, Regina Rudawska, Josep M. Trigo-Rodriguez, and
Jun-Ichi Watanabe.
In the following sections, we give some information
about the MDC structure, about the meteor shower data
included in the database, and we discuss some problems
October 3, 2018
encountered in the archiving process. Finally, we address
to some future plans related with MDC.
2. Content of the lists of meteor showers
At the time of the Meteoroids 2013 meeting in Pozna´n
the IAU MDC database stored the data of 579 showers
(Jopek and Kanuchov´a, 2014). The list of established
meteor showers contained 95 records, the working list
included 460 meteor showers, among them 95 had pro
tempore status. The list of shower groups contained 24
complexes, three of them had the established status. At
present 700 showers are archived and their parameters are
grouped into five listings.
1. List of all showers. It contains actually 700 show-
ers. These are all showers (except the removed ones)
registered in the Database.
2. List of established showers contains 112 ones. A
shower can be moved from the Working list to the list
of established showers if: 1) it passes the verification
procedure and 2) is given an official name which has
to be approved by the IAU F1 Commission business
meeting during the next GA IAU.
Exceptionally, the established meteor shower may be
moved back to the working list (see section 3).
In the time period between the last and the next GA
IAU one can nominate a shower from the working
list to become an established shower (its IAU nu-
merical code appears with a green background in the
list). The list of nominated candidates is discussed
and completed shortly before the GA IAU and it is
proposed for the approval by the adequate Commis-
sion during the GA IAU.
3. The Working list contains 563 showers that were
already published in the scientific literature or in the
IMO WGN Journal and 36 new submissions with the
"pro tempore" flag (the IAU numerical code is with
the orange background). The "pro tempore" flag is
removed after publication (or acceptation for publi-
cation). The working list showers can stay on this list
for many years, until they fulfill verification criteria
and obtain official names accepted at the GA IAU.
If a new submission is not be published within 2
years, it is removed permanently from the database.
A shower can be removed from the database also on
the basis of the published analysis.
4. List of shower groups contains actually 25 shower
complexes. Three of them have the status "estab-
lished", one group obtained "pro tempore" flag.
5. List of removed showers. In the past some showers
were included in the working list but for some reasons
they were removed from it. Because these showers
can be found in literature, we think that we should
keep complete information about the contents of the
database. The List of Removed Showers is a kind
of historical archive. Actually 10 showers are in this
list. They were removed from the Working list for
different reasons. In most cases a given shower had
been proved to be already observed and registered in
MDC.
The content of all lists mentioned above may be displayed
by the Web browser, four of them can be downloaded as
ASCII files.
As mentioned above, the list of all showers contains the
data on 700 showers. However, 228 of them are represented
by two or more sets of radiants and orbital parameters.
Additional sets of parameters were collected by us from
the meteor literature, but some of them originate from re-
cent observations and were submitted by the MDC users.
The shower MDC database contains inclusively 1184 data
records. Multiple sets of shower parameters play impor-
tant functions, apart from their cognitive importance, each
independent set of parameters strongly confirms the exis-
tence of a particular meteor shower.
3. Newly established showers
Meteor showers were officially named for the first time
in the history of astronomy at the XXVIIth GA IAU held
in Rio de Janeiro in August 2009. At this conference 64
meteor showers were given official names (Jopek and Jen-
niskens, 2011). At the following XXVIIIth GA IAU in
Beijing 2012 further 31 official names of the meteor show-
ers were approved (Jopek and Kanuchov´a, 2014). Other
18 showers were named officially in 2015 at the XXIXth
GA IAU in Honolulu. Their geocentric and heliocentric
parameters are given in Table 1.
Thirteen of the showers that have been named recently
have radiants located on the northern hemisphere, most
likely because of the distribution of meteor observing sites
on the Earth. About half of the newly established show-
ers moves on prograde orbits. Only five showers have near
ecliptic orbits. Particles of eleven showers collide with the
Earth at the geocentric velocities greater than 50 km/s.
The shower 510/JRC June rho Cygnids has an exceptional
mean orbit that is perpendicular to the plane of the eclip-
tic.
At the GA IAU in Honolulu, at the request of Peter
Jenniskens, the Working Group on Meteor Shower Nomen-
clature proposed to move the established meteor shower
3/SIA (the Southern iota Aquariids) back to the working
list (for more details see Jenniskens et al., 2015).
Iden-
tification of a reliable meteor shower is a quite complex
task. The new results are confronted with those obtained
in the past, which, due to the complex structure of the
meteoroid streams and their variable dynamical evolution,
hinders finding a final solution. Therefore, switching of
the meteor shower status should be considered as some-
thing inevitable. Fortunately, such changes happen only
exceptionally.
2
Table 1: Geocentric radiants and heliocentric orbital data of 18 meteor showers (streams) officially named pending the XXIXth GA IAU
held in Honolulu in 2015. The solar ecliptic longitude λS at the time of shower maximum activity, the geocentric radiant right ascension and
declination αg, δg and the values of the angular orbital elements ω, Ω, i are given for the epoch J2000.0. Numerical values of the meteoroid
parameters are the average values determined from N individual sets of radiants and orbits.
IAU
Shower name
No Code
alpha Virginids
AVB
21
SSG
69
Southern mu Sagittariids
NCC Northern delta Cancrids
96
97
SCC
Southern delta Cancrids
h Virginids
HVI
343
362
JMC
June mu Cassiopeiids
428 DSV December sigma Virginids
431
506
510
512
524
526
529
530
July xi Arietids
533
549
49 Andromedids
569 OHY omicron Hydrids
JIP
FEV
JRC
RPU rho Puppids
LUM lambda Ursae Majorids
SLD
EHY eta Hydrids
ECV eta Corvids
JXA
FAN
June iota Pegasids
February epsilon Virginids
June rho Cygnids
Southern lambda Draconids
λS
deg
32.0
86.0
296.0
296.0
38.0
74.0
262.0
94.0
314.0
84
231.0
215.0
221.6
256.9
302
119
112.0
309
αg
deg
203.5
273.2
127.6
125.0
204.8
17.5
200.8
332.1
200.4
321.8
130.4
158.2
163
132.9
192.0
40.1
20.5
176.3
δg
deg
2.9
-29.5
21.5
14.4
-11.5
53.9
5.8
29.1
11.0
43.9
-26.3
49.4
68.1
2.3
-18.1
10.6
46.6
-34.1
Vg
km/s
18.8
25.1
27.2
27.0
17.2
43.6
66.2
58.5
62.9
50.2
57.8
60.3
48.7
62.5
68.1
69.4
60.2
59.1
a
au
2.55
2.02
2.23
2.26
2.28
57.24
8.18
7.44
8.28
21
9.40
13
4.0
15
5.29
-
7.71
-
q
au
0.744
0.457
0.410
0.430
0.742
0.577
0.565
0.903
0.491
1.007
0.987
0.917
0.986
0.383
0.820
0.883
0.898
0.684
e
0.716
0.769
0.814
0.811
0.659
0.990
0.971
0.928
0.954
0.931
0.915
0.931
0.744
0.974
0.847
0.965
0.922
0.931
ω
deg
274.9
104.5
286.6
105.0
72.7
97.68
97.9
219.9
272.5
190
349.4
147
189
103.8
50.1
318
139.8
68.6
Ω
deg
30.0
266.4
290.0
109.3
218.2
74.0
261.8
94.1
312.6
84.2
50.8
215.0
221.6
76.9
122.2
299
118.0
128.9
i
deg
7.0
6.0
2.7
4.7
0.9
68.3
151.5
112.8
138.0
90
107.0
115
88.0
142.8
50.1
171.6
117.9
114.3
N
12
70
74
69
11
584
22
11
55
16
22
29
26
120
16
61
76
29
4. Modification of the meteor shower nomencla-
5. Errors, mistakes, shortcomings and positives
ture rules
The traditional meteor shower nomenclature was for-
malized by adopting a set of rules (Jenniskens, 2007, 2008;
Jopek and Jenniskens, 2011). The basic rules predicated
"The general rule is that a meteor shower should be named
after the constellation of stars that contains the radiant
. . . " and also "to distinguish among showers from the same
constellation the shower may be named after the nearest
(brightest) star . . . ". However, sometimes such definitions
are not sufficient from the practical point of view. One
can see in Fig. 1 that the radiant point of January nu
Hydrids is placed very close to the border between the
Hydra and Sextans constellations. With "naked eye" it
is not possible to decide which constellation this radiant
lies in. In another case the radiant clearly lies in a given
constellation, but the nearest stars belongs to another one,
see Fig. 1. For these reasons, the current wording in the
nomenclature rule reading: "The general rule is that a
meteor shower (and a meteoroid stream) should be named
after the current constellation that contains the radiant,
specifically using the possessive Latin form" should be re-
placed with a new rule reading: "The general rule is that a
meteor shower (and a meteoroid stream) should be named
after the constellation that contains the nearest star to
the radiant point, using the possessive Latin form". This
nomenclature change was adopted during the IAU Com-
mission 22 business meeting in Helsinki at the ACM 2014
conference. It was clarified in the discussion that the near-
est star with a Bayer designation, a Greek or Roman let-
ter, or (in exceptional cases) Flamsteed number is meant.
This nomenclature change was approved by voting at the
XXIXth IAU GA in Honolulu.
3
The MDC shower database is certainly not perfect -- it
contains some erroneous data, mistakes and various short-
comings. The reasons for this state are many varying from
our distraction to insufficient experience. However nobody
is perfect and the erroneous data can be also found in the
literature from which we derived the meteor shower data.
The meteor parameters given in the MDC are not ho-
mogeneous. They originate from different epochs and they
are determined from visual, photographic, video and radar
observations. Thus the uncertainties of the parameters are
different. There is another problem (also noticed by An-
drei´c et al. (2014)). Some authors supply to the MDC
only geocentric parameters without orbital information.
Furthermore, sometimes apparent radiant parameters are
submitted instead of the geocentric ones. It is not always
possible to recognize the difference reading the scientific
source literature. Similarly, reading literature, quite often
one cannot find any information about the method used
for averaging of the meteor shower parameters. Different
methods result in different mean values of the radiant and
the orbital parameters. Thus, often inconsistencies are
found between the values of the semi-major axis, eccen-
tricity, perihelion distance and the argument of perihelion
of the meteoroid stream orbit given in MDC.
Hence, despite our efforts, MDC contains the erroneous
data and mistakes. Therefore with a view to improve the
database we really appreciate every critical remark related
to the information archived in MDC.
The MDC database is maintained on a voluntary basis,
which limits our ability to handle complex submissions and
to develop user interface utilities.
We appreciate very much all positive signals about
Figure 1: Top panel: the difficult case of the radiant placement
(marked by a star symbol) of January nu Hydrids (544/JNH). The
radiant lies almost at the border of two constellations. Bottom panel:
The radiant of the May lambda Virginids (148/MLV) lies in the Cen-
taurus constellation, but the closest star lies in the Hydra constella-
tion. This shower was named before any shower nomenclature rules
were formalized.
Figure 2: The radiant of the meteor shower 536/FSO, 47 Ophi-
uchids, (marked by a star symbol) plotted on two star maps: at the
top panel, the stars were taken from the subset of the PPM star cat-
alogue; and at the bottom panel, the stars originate from the BSC
catalogue. The nearest star to the radiant is in both cases marked
by a blue circle. As one can see, adjudication of the name of this
shower depends on the selection of the catalogue.
the MDC shower database, e.g.
"You might be in-
terested to know that colleagues and I are using the
information you sent me to estimate the location and
intensity of meteoroid streams striking the lunar sur-
face.
This will be used to help with the inter-
pretation of observations from the upcoming LADEE
mission: http://www.nasa.gov/mission_pages/LADEE/
main/index.html#.Ue1uT43VCSo.", (Stubbs, 2013). We
found the other two in Andrei´c et al. (2014): "The most
valuable option of the IAU MDC web page is the possi-
bility to download the current shower list . . . " and "Nice
touch is that the shower data frequently include the link
to the reference from which the data originate". We are
very grateful for all such encouraging opinions.
6. Further MDC shower database activity
At first, we have to continue our main task which is
the assignment of IAU codes to the new meteor showers
submitted to MDC.
As far as possible, the MDC shower database will be im-
proved. As first we will complete archiving of the original
source meteor data found in literature. Also we will com-
plete collecting the hyper-links to the ADS meteor shower
literature references.
Following the remark given by Andrei´c et al. (2014),
we will add the periods of activity of the meteor showers
to the database. We also plane to implement a software
which will ease the employment of the MDC database.
Despite the recent modification of the meteor shower
4
Lindblad B.A., 1991, in Origin and Evolution of Interplanetary Dust,
Proc. IAU Colloq. 126, Eds: Levasseur-Regourd and Hasegawa H.,
Kluwer Dordrecht, 311
Lindblad B.A., Steel D.I., 1994, in Asteroids, Comets, Meteors 1993,
Proc. IAU Symp. 16
Lindblad B.A., Neslusan L.,Porubcan V., Svoren J., 2003, EM&P,
93, 249
Neslusan L., Porubcan V., Svoren J., 2014, EM&P, 111, 105
Porubcan V., Svoren J., Neslusan L., Schunov´a E., 2011, Proceedings
of the Meteroids Conference held in Breckenridge, Colorado, USA,
May 24-28, 2010. Edited by W.J. Cooke, D.E. Moser, B.F. Hardin,
and D. Janches. NASA/CP-2011-216469., 338
Roeser S., Bastian U., 1991, The PPM Star Catalog, Astronomisches
Rechen-Institut, Heidelberg, Vols. I and II, printed by Spektrum
Akademischer Verlag, Heidelberg
Stubbs T., 2013, Private comunication
Svoren J., Porubcan V., Neslusan, 2008, EM&P, 102, 11
nomenclature rules, we suggest to the Working Group on
Meteor Shower Nomenclature to accept the Yale Bright
Star Catalogue (BSC), 5th Revised Ed. (Hoffleit and War-
ren, 1991) as the standard for the naming of the meteor
showers.
The reason for such standardization is clearly seen in
Fig. 2. The top panel was plotted using the subset of ∼ 900
stars taken from the Positions and Proper Motions Star
Catalogue (Roeser and Bastian, 1991). The bottom map
is plotted using stars from the Yale Bright Star Catalogue
containing 3141 stars. As shown adjudication of a shower
name depends on the choice of the nearest star to the
meteor radiant and thus it depends on the selection of
stars in the star catalogue.
Finally, we plan to move more showers to the List of
Removed Showers if such recommendation is published in
literature.
Acknowledgments
We would like to acknowledge the users of the IAU MDC
who informed us about several erroneous data: Zeljko An-
drei´c, Regina Rudawska and Damir Segon. Zeljko An-
drei´c, Damir Segon and Denis Vida are acknowledged for
their excellent "A statistical walk through the IAU MDC
database" (Andrei´c et al., 2014). Z.K. was supported
by VEGA - The Slovak Agency for Science, Grant No.
2/0032/14.
The authors acknowledge both anonymous referees for
their useful comments.
This research has made use of NASA's Astrophysics
Data System Bibliographic Services.
References
References
Andrei´c Z., Segon D., Vida D., 2014, Proceedings of the International
Meteor Conference, Giron, France, 18-21 September 2014, Eds.:
Rault, J.-L.; Roggemans, P., International Meteor Organization,
pp. 126 -- 133
Hoffleit D., Warren Jr., W.H., 1991, "The Bright Star Catalog,
5th Revised Edition (Preliminary Version)", http://tdc-www.
harvard.edu/catalogs/bsc5.html
Jenniskens P., 2007, Div. III, Comm.22, WG Task Group for Me-
teor Shower Nomenclature, IAU Information Bulletin 99, January
2007, 60
Jenniskens P., 2008, EM&P, 102, 5
Jenniskens, P., Borovicka, J., Watanabe J., Consolmagno, G., Jopek,
T.J., Vaubaillon, J., Abe, S., Janches, D., Ryabova, G., Ishiguro,
M., Zhu, J., 2015 Transactions IAU, Volume 10, Issue T28, 120
Jopek T.J., Jenniskens P.M., 2011, in Proc. Meteoroids 2010 Conf.,
held in Breckenridge, Colorado USA, May 24-28 2010, eds Cooke
W.J., Moser D.E., Hardin B.F., Janches D., NASA/CP-2011-
216469, p. 7
Jopek T.J., Kanuchov´a Z., 2014, Meteoroids 2013, Proceedings of the
Astronomical Conference held at A.M. University, Pozna, Poland,
Aug. 26-30, 2013, Eds.: T.J. Jopek, F.J.M. Rietmeijer, J. Watan-
abe, I.P. Williams, A.M. University Press, 2014, 353
Lindblad B.A., 1987,
in Interplanetary Matter, Eds: Ceplecha
Z., Pecina P., Publications of the Astronomical Institute of the
Czechoslovak Academy of Sciences, No. 67, 201
5
|
1912.10939 | 1 | 1912 | 2019-12-23T15:58:01 | The Detectability and Constraints of Biosignature Gases in the Near & Mid-Infrared from Transit Transmission Spectroscopy | [
"astro-ph.EP",
"astro-ph.IM"
] | The James Webb Space Telescope (JWST) is expected to revolutionize our understanding of Jovian worlds over the coming decade. However, as we push towards characterizing cooler, smaller, "terrestrial-like" planets, dedicated next-generation facilities will be required to tease out the small spectral signatures indicative of biological activity. Here, we evaluate the feasibility of determining atmospheric properties, from near-to-mid-infrared transmission spectra, of transiting temperate terrestrial M-dwarf companions. Specifically, we utilize atmospheric retrievals to explore the trade space between spectral resolution, wavelength coverage, and signal-to-noise on our ability to both detect molecular species and constrain their abundances. We find that increasing spectral resolution beyond R=100 for near-infrared wavelengths, shorter than 5$\mu$m, proves to reduce the degeneracy between spectral features of different molecules and thus greatly benefits the abundance constraints. However, this benefit is greatly diminished beyond 5$\mu$m as any overlap between broad features in the mid-infrared does not deconvolve with higher resolutions. Additionally, our findings revealed that the inclusion of features beyond 11$\mu$m did not meaningfully improve the detection significance nor abundance constraints results. We conclude that an instrument with continuous wavelength coverage from $\sim$2-11$\mu$m, spectral resolution of R$\simeq$50-300, and a 25m$^2$ collecting area, would be capable of detecting H$_2$O, CO$_2$, CH$_4$, O$_3$, and N$_2$O in the atmosphere of an Earth-analog transiting a M-dwarf (mag$_{K}=8.0$) within 50 transits, and obtain better than an order-of-magnitude constraint on each of their abundances. | astro-ph.EP | astro-ph | Draft version December 24, 2019
Typeset using LATEX twocolumn style in AASTeX62
The Detectability and Constraints of Biosignature Gases
in the Near & Mid-Infrared from Transit Transmission Spectroscopy
L. Tremblay,1 M.R. Line,1 K. Stevenson,2 T. Kataria,3 R.T. Zellem,3 J.J. Fortney,4 and C. Morley5
1School of Earth and Space Exploration, Arizona State University,
PO Box 871404, Tempe, AZ 85287-1404, USA
2Space Telescope Science Institute,
3700 San Martin Drive, Baltimore, MD 21218, USA
3Jet Propulsion Laboratory, California Institute of Technology,
4800 Oak Grove Drive, Pasadena, CA 91109, USA
4Department of Astronomy & Astrophysics, University of California Santa Cruz,
5Department of Astronomy, University of Texas at Austin,
Santa Cruz, CA, USA
Austin, TX, USA
ABSTRACT
The James Webb Space Telescope (JWST ) is expected to revolutionize our understanding of Jo-
vian worlds over the coming decade. However, as we push towards characterizing cooler, smaller,
"terrestrial-like" planets, dedicated next-generation facilities will be required to tease out the small
spectral signatures indicative of biological activity. Here, we evaluate the feasibility of determining
atmospheric properties, from near-to-mid-infrared transmission spectra, of transiting temperate ter-
restrial M-dwarf companions. Specifically, we utilize atmospheric retrievals to explore the trade space
between spectral resolution, wavelength coverage, and signal-to-noise on our ability to both detect
molecular species and constrain their abundances. We find that increasing spectral resolution beyond
R=100 for near-infrared wavelengths, shorter than 5µm, proves to reduce the degeneracy between spec-
tral features of different molecules and thus greatly benefits the abundance constraints. However, this
benefit is greatly diminished beyond 5µm as any overlap between broad features in the mid-infrared
does not deconvolve with higher resolutions. Additionally, our findings revealed that the inclusion of
features beyond 11µm did not meaningfully improve the detection significance nor abundance con-
straints results. We conclude that an instrument with continuous wavelength coverage from ∼2-11µm,
spectral resolution of R(cid:39)50-300, and a 25m2 collecting area, would be capable of detecting H2O, CO2,
CH4, O3, and N2O in the atmosphere of an Earth-analog transiting a M-dwarf (magK = 8.0) within
50 transits, and obtain better than an order-of-magnitude constraint on each of their abundances.
Keywords: terrestrial exoplanets, atmospheric retrievals, mid-infrared, biosignatures
1. INTRODUCTION
Characterizing the climate and composition of terres-
trial atmospheres is a primary goal of exoplanet science
over the next decades (NAS 2020 Vision1). Temperate
terrestrial exoplanets are of course, of great interest due
to their possible astrobiological implications. Namely,
their potential to develop and maintain volatile-rich sec-
ondary atmospheres capable of supporting biology. As
Earth is the only known planet to host life, it is straight-
1
https://sites.nationalacademies.org/cs/groups/bpasite/
documents/webpage/bpa 064932.pdf
forward that we would begin our search for extraterres-
trial life by considering the conditions on Earth as the
primary conditions which must exist for the develop-
ment of life as we know it.
The principal gases that comprise Earth's atmosphere
(nitrogen/nitrous oxide, water, carbon dioxide, oxy-
gen/ozone, and methane) are generally regarded as "bio-
indicators" meaning that their presence is indicative of
potentially habitable atmospheric conditions. "Biosig-
natures" are those observables that are representative
of biological processes affecting the planet's atmospheric
chemistry (i.e. the presence of life). Catling & Kasting
(2007) showed that all of the bulk gases in Earth's at-
9
1
0
2
c
e
D
3
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
3
9
0
1
.
2
1
9
1
:
v
i
X
r
a
2
Tremblay et al.
mosphere, excluding inert gases, are influenced by bio-
genic processes. Life on Earth is the primary contrib-
utor to the global chemical disequilibrium (Krissansen-
Totton et al. 2016), and for this reason disequilibrium
is presumed to be a robust biosignature (Cockell et al.
2009; Kasting 2009; L´eger 2000; Sagan et al. 1993; Sea-
ger 2014; Seager & Bains 2015; Seager & Deming 2010).
The combinations of molecules responsible for Earth's
disequilibrium are O3+CH4 and O3+N2O (Krissansen-
Totton et al. 2016). If detected at certain abundances,
the presence of these molecules simultaneously cannot
be plausibly explained by abiotic processes (Domagal-
Goldman et al. 2014; Harman et al. 2015, 2018; L´eger
et al. 1993; Meadows et al. 2018; Meadows 2017; Tian
et al. 2014; Schwieterman et al. 2016). As such, a key
goal of exoplanet exploration is to quantify the presence
of these gases on nearby terrestrial worlds.
Due to the favorable planet-to-star radius ratio and
occurrence rates, M-dwarf systems offer the best near-
term opportunity for characterizing temperate terres-
trial worlds and the possible presence of biosignatures.
Ground-based radial velocity surveys and TESS are ex-
pected to detect dozens of temperate terrestrial-sized
(0.8 - 1.6R⊕) exoplanets around M-dwarf's (Quirren-
bach et al. 2018; Barclay et al. 2018; Kempton et al.
2018; Zechmeister et al. 2019), potentially amenable to
follow-up with JWST. Numerous previous works have
demonstrated that JWST will likely provide astound-
ing constraints on the atmospheric properties of Jovian-
to-Super-Earth planets (Deming et al. 2009; Beichman
et al. 2014; Barstow et al. 2014; Greene et al. 2016;
Batalha et al. 2018; Rocchetto et al. 2016; Benneke &
Seager 2012). However, it remains an open question as
to how well JWST will perform as it is pushed to ob-
serve temperate terrestrial bodies with smaller spectral
features. Possible limitations may stem from a com-
bination of large detector noise floors (similar detector
technology as HST WFC3 and Spitzer IRAC, Greene
et al. (2016)), saturation limits for the brightest tar-
gets (Batalha et al. 2018), and lack of continuous near-
to-mid-infrared wavelength coverage in a single mode
(greatly increasing the number of required transits).
The discovery of the TRAPPIST-1 system (Gillon
et al. 2016) has lead to a number of works investigat-
ing the climate, composition, and internal structure of
temperate M-dwarf worlds (Unterborn & Panero 2017;
Turbet et al. 2018; Suissa & Kipping 2018; Meadows
2017; Wunderlich et al. 2019). Complementary to these
efforts, recent studies have explored the capabilities of
JWST to characterize various atmospheric compositions
on TRAPPIST-1 planets through both transmission and
emission spectroscopy (Barstow & Irwin 2016; Morley
et al. 2017; Batalha et al. 2018; Krissansen-Totton et al.
2018; Lustig-Yaeger et al. 2019; Wunderlich et al. 2019).
Morley et al. (2017) concluded that JWST 's NIR-
Spec G235M/F170LP mode (1.66 - 3.07µm) would be
capable of rejecting a flat-line spectrum at 5σ confi-
dence within 20 transits for a range of atmospheric
compositions. While a flat-line rejection test is valu-
able,
it does not guarantee the ability to detect the
presence of any specific molecule with high statistical
confidence.Lustig-Yaeger et al. (2019) concluded that
transmission spectroscopy with NIRSpec PRISM is op-
timal for detecting the presence of high mean molec-
ular weight atmospheres within ∼12 transits. Addi-
tionally, they find that CO2,
if present, will be eas-
ily detectable regardless of the atmospheric composi-
tion and MIRI-LRS may be capable of detecting the
9.6µm O3 on planet 1e (with an SNR=3) with greater
than 100 transits. Wunderlich et al. (2019) produced
self-consistent forward models for the atmosphere of an
Earth-like planet around early-to-late M-dwarfs. Their
"Earth around TRAPPIST-1" model notably predicts a
greatly increased temperature in the mid-to-upper at-
mosphere, inflating the scale height and increasing the
detectability of molecular species for transmission spec-
troscopy with JWST. Furthermore, the predicted chemi-
cal evolution indicated a significant enhancement of CH4
and H2O compared to modern Earth and subsequently,
that their corresponding spectral features, along with
CO2, would be detectable within ∼10 transits when us-
ing the appropriate NIRSpec high-resolution filters.
Barstow & Irwin (2016) performed a retrieval analy-
sis on synthetic spectra of the inner-most TRAPPIST-1
companions (b, c, d) each with an Earth-like compo-
sition atmosphere. Utilizing a combination of NIR-
Spec PRISM and MIRI-LRS, they conclude that O3, at
Earth-like abundances, would be detectable on planets
1c and 1d with 30 transits on each instrument. Also
utilizing an atmospheric retrieval analysis, Krissansen-
Totton et al. (2018) analyzed the detectability of the
CO2+CH4 disequilibrium biosignature, suspected to
have been present in Earth's early atmosphere when
the abundances of each were significantly enhanced
compared to modern-day values.
They concluded
that JWST would be capable of obtaining order-of-
magnitude constraints on the mixing ratios of both CH4
and CO2 in an Archean Earth-like atmosphere, in ∼10
transits. Alternatively, the abundance of O3 in a mod-
ern Earth-like atmosphere could not be constrained by
NIRSpec Prism and would be completely unbounded by
MIRI-LRS.
While it may be possible for JWST to detect and con-
strain the dominant molecular species in certain terres-
trial atmospheres with advantageous compositions, ex-
isting literature does not substantiate its capabilities to
detect or constrain the abundances of the five primary
bio-indicators in a modern Earth-like atmosphere on an
M-dwarf companion.
The primary aim of this work is to develop a base-
line understanding of how key observational parameters
(spectral resolution, wavelength coverage, and signal-to-
noise) of near-to-mid-infrared transmission spectra in-
fluence the ability to detect and constrain biologically
relevant molecular species in the atmospheres of transit-
ing Earth-like planets orbiting M-dwarfs. We leverage
powerful Bayesian retrieval tools to obtain both parame-
ter constraints and Bayesian molecular detection signif-
icances as a function of the instrumental trades. In §2
we describe our model/retrieval and instrumental-trade
space setup, §3 summarizes the key results, followed by
a discussion and key points in §4 and §5.
2. SIMULATION SET UP
Here, we outline the development of our forward
model used to compute the synthetic transmission spec-
tra, elaborate on the instrumental trade space, discuss
the details of our instrument noise model, and review
our retrieval approach.
2.1. Transmission Forward Model
We simulate the transmission spectrum of a terrestrial
planet, with mass and radius equivalent to TRAPPIST-
1e, residing within the "habitable zone" of TRAPPIST-1
with an approximate modern Earth-like atmosphere (ob-
servational system setup/signal-to-noise discussed be-
low). We acknowledge, up front, that this may not
be a physically plausible scenario due to the vastly dif-
ferent incident spectral energy distribution. However,
given the overwhelming number of unknowns involved
in self-consistent planetary atmosphere modeling (e.g.,
star-planet interactions, surface fluxes, formation condi-
tions, bulk elemental composition, 3D atmospheric dy-
namical effects, cloud micro-physics, dynamical evolu-
tion/history) we simply choose to treat our atmosphere
as "Earth-like" (as has also been done in previous works,
Morley et al. (2017); Krissansen-Totton et al. (2018)).
To produce model transmission spectra, we leverage a
variant2 of the CHIMERA3 transmission spectrum routine
(Batalha & Line 2017; Line et al. 2013; Greene et al.
2016; Line & Parmentier 2016) (Batalha & Line 2017;
Line et al. 2013; Greene et al. 2016; Line & Parmentier
3
Figure 1. Validation of our transmission forward model
(at R=100) against the Virtual Planetary Laboratory (VPL)
Earth model3 utilizing the same temperature-pressure and
gas mixing ratio profiles. The spectra were scaled to a 1R⊕
around a 0.15R(cid:12) star. For reference, overlaid are 5 ppm
error bars.
2016). Specifically, we re-parameterize the code (Table
1) to make it more amenable for temperate worlds by
including as free parameters the constant-with-altitude
mixing ratios of H2O, CO2, CO, CH4, O3, N2O and
an "unknown" background gas with a free-parameter
molecular weight (taken to be earth's N2+O2 value),
an isothermal "scale-height" temperature, planetary ra-
dius at the surface (or scaling there-of), and an opaque
gray cloud-top-pressure (nominal truth values given in
Table 1). We mimic refraction with this cloud top pres-
sure (here, 0.56 bars) using the prescription from Robin-
son et al. (2017), but otherwise assume a cloud free
atmosphere (the refractive layer is in affect, mimick-
ing a cloud, as well as its inclusion as a free parame-
ter). CHIMERA uses correlated-K opacities (computed at
the given constant resolving powers below), here derived
from a grid of pre-computed line-by-line (≤0.01 cm−1,
70 - 410 K, 10−7 - 30 bar) cross-sections generated with
the HITRAN HAPI Routine (Kochanov et al. 2016) and
the HITRAN2016 line database (Gordon et al. 2017).
We do not include collision induced opacities, though
they may spectrally present themselves throughout the
mid-infrared (Schwieterman et al. 2015).
We benchmarked our transmission forward model
against those available from the Virtual Planetary Lab-
oratory4 (Figure 1).
2
https://www.dropbox.com/sh/nehnyq6mlkfb13w/
AABuDn-uFqTvAptWVSbGoA04a?dl=0
3 https://github.com/mrline/CHIMERA
4
http://depts.washington.edu/naivpl/content/
vpl-spectral-explorer)
4
Tremblay et al.
Model Value Uniform Prior
log(H2O)
log(CO2)
log(CH4)
log(O3)
log(N2O)
log(CO)
Tiso
xRP
log(CTP)
Bkg MMW
-5.5
-3.45
-6.3
-6.5
-6.3
-7.0
280K
0.918R⊕
-0.25
28.6
[-12, 0]
[-12, 0]
[-12, 0]
[-12, 0]
[-12, 0]
[-12, 0]
[100, 800]
[0.5, 1.5]
[-6, 2]
[2.0, 44.0]
Table 1. The 10 free parameters, the model values, and
their associated uniform priors used in our CHIMERA radiative
transfer and retrieval code.
2.2. Parameter Estimation & Model Selection
Approach
We perform Bayesian parameter estimation and model
selection, on the simulated datasets described below,
using the PyMultiNest routine (Buchner et al. 2014)
following the methods described in Benneke & Seager
(2013). We initialized our retrievals with 3000 live
points. An advantage of nested sampling algorithms is
ease of evidence computation, which can be used to as-
sess model complexity. We use Bayesian nested model
comparison (by removing one gas at a time) to deter-
mine the detection significance of each molecular species
(Trotta 2008; Benneke & Seager 2013), and is what we
utilize as a metric for assessing instrument performance.
An advantage of utilizing the detection significance via
the Bayesian evidence, over more traditional methods
(e.g., line or band height above a continuum relative to
the noise) is that it fully utilizes all of the spectral in-
formation included in all of the relevant bands as well
as encompasses the model degeneracies.
2.3. Simulated Spectrograph & Radiometric Noise
Model
We use the JWST radiometric noise model described
in Greene et al. (2016) to produce spectrophotometric
uncertainties. The collecting area of 25m2 was retained
to reflect the identical collecting area of JWST in or-
der to provide a baseline with which to easily scale the
results of this study. Modifications were made to param-
eters of the code to reflect the expected performance of
a next-generation mid-infrared telescope. These param-
eters included a 2.85 - 30.0 micron wavelength range
(HgCdTe detectors below 10.5 microns and SiAs detec-
tors above 10.5 microns), a zodiacal background esti-
mated by Glass & Blair (2015); Glass et al. (2015), and
an intrinsic resolving power of 300. These modifications
reflect the capabilities of new detector technologies dis-
cussed by Matsuo et al. (2018).
Our input stellar spectrum derives from the PHOENIX
stellar models using the PySynPhot software pack-
age, Lim et al. (2015). We adopted the values for
TRAPPIST-1 (T=2550K, M/H=0.40, log(g)=4.0) and
stellar radius of 81373.5km from Gillon et al. (2016).
Additionally, we have scaled the stellar spectrum to a
K-band magnitude of 8.0 rather than magK = 10.3,
TRAPPIST-1's native magnitude. While this may be
representative of an optimistic scenario, magK = 8.0
is the mean magnitude of the brightest 10 M-dwarf
stars in the Barclay catalog of the predicted TESS yield
(Barclay et al. 2018).
Short Wavelength Long Wavelength Spectral Resolution
Boundary
Boundary
1µm
3µm
5µm
5µm
11µm
30µm
5µm
11µm
30µm
11µm
30µm
300
100
50
100
50
100
50
30
300
100
50
300
100
50
100
50
30
100
50
30
100
50
30
Table 2. A breakdown of the 23 test cases evaluated in this
study as a combination of wavelength coverage and spectral
resolution.
2.4. Observational Parameter Space
We produce synthetic transmission spectra of our
Earth-like atmosphere model over a grid of spectral res-
olution and bandpass choices. The nominal wavelength
coverage was chosen to cover the near and mid-infrared
5
Figure 2. Wavelength region explored in this work (here, shown at R=50). The colored bands indicate the presence, and
approximate width, of a spectral feature for a given molecule. We explore constraints obtainable over several wavelength bands
within this region (in Table 2).
Figure 3. Influence of resolution on spectral features. Lower resolution results in the blending of features, causing a loss of
information. The vertical lines at 3, 5, and 11µm denote the division of the wavelength regimes explored in this study.
H2OCO2CH4N2OO3COH2OCO2CH4N2OO3COCO2CO2CH4H2OH2OCO2H2ON2OCH4O3O3N2OH2ORRRes = 300Res = 30Res = 50Res = 1006
Tremblay et al.
regimes spanning 1-30µm, shown in Figure 2. We divide
this wavelength range into 8 separate regimes to explore
the influence of additional bands of a given molecule
and their overlap with features of other molecules. Ad-
ditionally, we test low and moderate spectral resolutions
ranging from R=30 to R=300. The spectral resolution
(λ/δλ) dictates the degree to which individual molecu-
lar bands can be resolved and overlap degeneracies bro-
ken. Figure 3 compares the spectra at the resolutions
explored in this work. We simulate different resolutions
within each spectral band in order to explore the resolu-
tion trade along with each wavelength regime. This pro-
duces a grid of 23 test cases (shown in Table 2) to eval-
uate through our retrieval algorithm described in 2.2.
In addition to resolution and wavelength coverage, we
also explored the signal-to-noise trade, parameterized
here as the number of transits (1-100), which can be
considered a proxy for mirror diameter or source magni-
tude, where the noise on a single transit is defined by the
"nominal" noise model setup described in 2.3. Utilizing
this trade, we explored the minimum number of transits
necessary to achieve a threshold 3.6σ confidence level
detection of each molecule. We provide equations 1 &
2 as a means of converting our retrieval results to stars
of differing magnitudes or for different sized collecting
mirrors.
Mef f = Mnom − 2.5 log10(
Tef f
Tnom
)
(1)
Where Tef f /Tnom is the ratio determining how many
more or less transits are equivalent to a change from the
native magnitude (Mnom = 8.0) to a new magnitude
(Mef f ).
(cid:114) Tef f
Tnom
Aef f = Anom
(2)
Where Tef f /Tnom is the ratio of transits equivalent to a
change from the native collecting area (Anom = 25m2)
to a new mirror size (Aef f ).
3. RETRIEVAL RESULTS
For the majority of our test cases, the inclusion of
a broader bandpass decreases both the necessary ob-
servation time to claim a statistically significant detec-
tion and reduces the uncertainty on the abundance con-
straints. Therefore, with few molecule-specific excep-
tions, we find that altering the choice of wavelength cov-
erage has a more prominent benefit for both our detec-
tion significance values and abundance constraints than
the choice of spectral resolution (Tables 3 & 4).
Our detection significance analysis directly compares
each bandpass and resolution combination by determin-
ing the number of transits required to achieve a 3.6σ
detection for a specific molecule. Table 3 displays these
required number of transits for each molecule over each
combination of resolution and wavelength coverage.
Table 4 enumerates the 1σ error widths for each
molecule within each of our 23 test cases at 50 tran-
sits. Figures 4,6,7,9, and 10 within each of the following
sections reveal the influence that an increasing number
of transits has on the abundance constraints, for each
molecule individually at R=100. The following subsec-
tions provide an in-depth analysis of each molecule and
the most relevant transitions that influence our abun-
dance constraints and detection significance results.
3.1. Water (H2O)
Figure 4. Compares the improvements of the abundance
constraints for H2O, with increasing numbers of transits at
three resolutions for the two key bandpass choices.
Water has spectral features spanning nearly the entire
wavelength range in this study, with the most promi-
nent features centered at 2.65µm and 6.3µm. Addition-
ally, there are two smaller features extending into the
near-infrared at 1.4µm and 1.85µm as well as a series
of features upwards of 20µm. The most relevant com-
parison is between the near-infrared (which we define
as the 1-5µm bandpass) and the mid-infrared (5-11µm).
In isolation from other water features in the 5-11µm re-
gion, the broad 6.3µm feature produces comparatively
underwhelming results (for both detection significance
and abundance constraints) likely due to the limited
contribution from the few other prominent features in
this bandpass (Benneke & Seager 2012).
In contrast,
the 1-5µm region encompasses the prominent 2.65µm
feature as well as the two smaller features. The preci-
sion on the abundance constraints improves by approx-
imately 60% both when increasing the resolution from
R=50 to R=100 and again from R=100 to R=300. How-
ever, it is worth noting that these near-infrared features
are very narrow and are thus significantly more sensi-
R50R100R3001-5m3-11m7
Figure 5. Abundance constraints for each molecule within three key bandpasses (1-5µm, 3-11µm, and 5-30µm at R=100) as a
function of the number of observed transits. The heights of each histogram window has been fixed to a constant value to allow
for a direct comparison of the shape of the marginalized posterior probability distributions for each gas. The "hard edge" of
some distributions, near low abundances, indicates those that extend beyond the prior range.
5 Transits10 Transits25 Transits50 Transits100 Transits8
Tremblay et al.
tive to changes in the choice of spectral resolution. At a
resolution of R=100, a single H2O feature in the near-
infrared consists of between 10-20 resolution elements
compared to 44 for the 6.3µm feature.
While the near-infrared (1-5µm bandpass) produces
the tightest abundance constraints, it does not provide
the greatest detection significance values given the same
spectral resolution and number of transits (refer to Ta-
ble 3 for comparison of detection significance results).
By comparison, the 3-11µm range (which does not in-
clude additional H2O features) requires 33.9% less ob-
servation time at R=100 and 50.0% less at R=50 than
the 1-5µm bandpass. Alternatively, by looking further
into the mid-infrared, the 5-30µm bandpass offers a 3.4%
decrease at R=100 and a 24.5% decrease at R=50 com-
pared to the 1-5µm bandpass.
3.2. Carbon Dioxide (CO2)
(2016) derive the function for spectral modulation with
respect to wavelength for multiple absorbing molecular
species, reproduced here for convenience (Equation 3).
(cid:18) d ln(σλ,1)
dλ
(cid:19)
dαλ
dλ
=
2Rp
R2
(cid:63)
H
1
1 + ξ2σλ,2
ξ1σλ,1
+
ξ2σλ,2
ξ1σλ,1
d ln(σλ,2)
dλ
(3)
Where dαλ/dλ is the wavelength dependent slope of
the transmission spectra. H is the scale height given by
kbT
µg and σλ,i and ξi are the absorption cross section and
abundance, respectively, of a given molecular species.
In the case of CO2, the ξ1σλ,1 term is so much greater
than the other absorbers that the spectral modulation
due to CO2 becomes insensitive to even large changes
in the abundance. The retrieval thus determines that a
wide range of abundance values can reproduce the shape
of the observed features within the error bars. There-
fore, the only means to narrow the constraint is to in-
crease the number of observed transits.
At 50 transits, the 1-5µm bandpass produces approxi-
mately a single order-of-magnitude (1.01dex) constraint
at R=100 or 0.81dex at R=300. The abundance con-
straints produced by the 1-5µm bandpass cases at both
R=50 and R=100 are comparable to those of both the
3-11µm and 3-30µm bandpasses.
3.3. Methane (CH4)
Figure 6. Compares the improvements of the abundance
constraints for CO2, with increasing numbers of transits at
three resolutions for the two key bandpass choices.
Due to its prominent spectral features across most of
the near and mid-infrared, CO2 is the easiest molecule
to detect within each of our bandpass choices, with the
exception of 5-11µm. CO2 possess four prominent spec-
tral features centered at 2.0, 2.7, 4.3, and 15um. The
5-11µm region is the only bandpass choice which does
not include a spectral feature for CO2. Any other choice
of wavelength coverage and spectral resolution evaluated
in our study proved to be sufficient to detect CO2 in less
than 7 transits. The inclusion of the 4.3µm feature en-
sures a 3.6σ detection within the observation time of
approximately a single transit.
Despite the strong detection of CO2, the abundance
constraints are ironically, not as precise as would be ex-
pected. This is largely due to the fact that deriving
narrow abundance constraints relies on a change in the
strengths of the spectral features. Line & Parmentier
Figure 7. Compares the improvements of the abundance
constraints for CH4, with increasing numbers of transits at
three resolutions for the two key bandpass choices.
Due to its low abundance in the modern Earth at-
mosphere, CH4 is difficult to detect and ultimately de-
termines the lower limit on the observing time required
to detect all five bio-indicator molecules. Despite the
larger number of transits required to detect CH4, we
can place the best constraints on its abundance.
In-
deed, it is notable that our results indicate that for the
R50R100R3001-5m3-11mR50R100R3001-5m3-11m3-11µm at R=100 case, a 3.6σ detection is only possible
after 61.7 observed transits while at only 50 transits we
can place an order-of-magnitude constraint on its abun-
dance. This appears initially paradoxical, until we ex-
amine the underlying dependencies of these two different
results. Effectively, the detectability (via the Bayes fac-
tor) depends on the amplitude of the spectral features
with respect to the continuum or adjacent features of
other molecules (e.g., wavelength dependent derivative
of transit depth shown in equation 3). The mixing ra-
tio constraints on the other hand are dictated by the
derivative of the transit depth with respect to a given
species abundance (or rather, the inverse, e.g., Line et al.
(2012)), with each species having varying sensitivity.
Therefore, one could observe small features resulting in
a low detectability (as in the case of CH4) while having
a large abundance derivative, resulting in a tight con-
straint. Figure 8 illustrates the molecule-specific rela-
tionship between the abundance constraints (log) and
detection significance under the 1-30µm, R=100 sce-
nario for varying numbers of transits. Figure 8 illu-
minates the key differences between CH4 and CO2 in
that while CO2 has a high detection significance, it does
not necessarily have a precise constraint. Alternatively,
even low confidence detections of CH4 can yield precise
abundance constraints.
Figure 8. The relationship between detection significance
and abundance constraint. The numbers on each data point
represent the number of transits observed to achieve those
respective results. The number of transits for the N2O (or-
ange) are identical to those of H2O (blue), conveniently adja-
cent here. These results are representative of an instrument
with 1-30µm coverage, R=100 for a stellar mK =8 and mirror
aperture of 25m2). Unsurprisingly, an increase in detection
significance is correlated with higher precision. However,
the mapping between detection significance and constraint
is molecule specific.
At Earth-like abundances, CH4 has only three dis-
cernible spectral features in the near and mid-infrared:
2.3, 3.3, and 7.6µm. The most prominent of these is cen-
tered at 7.6µm and directly overlaps with a larger N2O
9
feature centered at 7.7µm. Bandpass choices that ex-
clude additional methane features will lack the ability to
overcome this degeneracy. Fortunately, the 3.3µm (ν3)
feature provides an unambiguous marker for the pres-
ence of CH4. We find that the inclusion of this feature is
crucial to detecting CH4 as neither the 5-11µm nor the
5-30µm bandpasses are capable of detecting methane
within 100 transits, despite the presence of the 7.6µm
(ν4) feature. There is an additional feature at 2.3µm
which, when combined with the 3.3µm band, greatly
improves the detection significance.
When observing only the 3-5µm bandpass, at R=300,
CH4 is detectable with 54.9 transits and provides an
abundance constraint of 1.04dex at 50 transits. Main-
taining this narrow wavelength coverage and reducing
to R=100, increases the necessary observation time to
beyond 100 transits. However, expanding our bandpass
either to the near or mid-infrared regimes proves to pro-
vide significant benefits. The near-infrared bandpass
(1-5µm) requires 38.5 transits and provides a 0.63dex
(at 50 transits) abundance constraint or 54.4 transits
and 0.81dex for R=300 and R=100, respectively. Alter-
natively, observing the 3-11µm bandpass requires 34.3
transits and provides a 0.79dex (at 50 transits) abun-
dance constraint or 61.7 transits and 0.99dex for R=300
and R=100, respectively.
Based on these results, we find that including the ν3
feature, within the chosen bandpass, is essential to de-
tecting and constraining CH4. In the scenario that spec-
tral resolution is limited to R=100, the addition of the
2.3µm feature is incredibly beneficial, providing a 0.5dex
improvement in the abundance constraint and requiring
only a third of the observational time. However, while
it is certainly ideal to include both features if possible,
it is quite noteworthy that the benefit of a higher res-
olution (R=300) choice, in the absence of the 2.3µm
feature, outweighs the benefit of including this feature
at a lower resolution (R=100).
3.4. Ozone (O3)
Ozone is different from the other molecular species
we have addressed thus far, in that its most prominent
features extend further into the mid-infrared. The three
primary features for O3 are at 4.75, 9.6, and 14.2µm with
the 9.6µm feature being the strongest. Conveniently,
due to the location of these spectral features and our
bandpass choices, we can evaluate the benefit of each
additional feature, individually, on the retrieval results.
Isolating the 9.6µm feature, in the 5-11µm bandpass,
requires a relatively low number of transits at only 14.9
and 18.2 transits for resolutions of R=100 and R=50,
respectively. However, the abundance constraints for
10
Tremblay et al.
1-5µm
1-11µm
1-30µm
38.5
54.4
-
44
52.1
43.8
50
-
-
-
-
8.5
9.4
8.3
9.1
10.9
Bandpass Resolution H2O CO2 CH4 O3 N2O
7.7
9.2
14.3
5.7
7.4
5.5
7.2
9.0
11.2
14.4
23.8
5.6
7.9
10.6
7.7
9.6
12.2
94.0
34.3
61.7
3-11µm
3-30µm
3-5µm
54.9
-
-
18.1 < 1
23.3 < 1
39.2 < 1
< 1
6.5
7.3
< 1
< 1
6.3
< 1
7.0
8.5
1.13
1.03
-
1.08
-
1.13
-
11.9 < 1
1.07
15.4
19.6
1.11
14.4 < 1
17.4 < 1
21.1
1.15
33.1
50
67.2
22.5
29.6
36.8
6.75
6.88
6.96
-
-
-
300
100
50
100
50
100
50
30
300
100
50
300
100
50
100
50
30
100
50
30
100
50
30
5-11µm
5-30µm
3-5µm
1-5µm
1-11µm
1-30µm
0.81
1.01
1.24
0.77
0.93
0.74
0.91
1.08
1.31
1.44
1.86
0.91
1.07
1.29
0.74
0.91
1.08
Bandpass Resolution H2O CO2 CH4 O3 N2O
0.77
0.97
1.19
0.74
0.92
0.72
0.89
1.07
1.16
1.27
1.65
0.85
1.02
1.23
0.99
1.21
1.38
1.70
2.01
2.03
1.50
1.69
1.78
0.65
↑
↑
0.60
0.75
0.59
0.73
0.87
0.95
↑
↑
0.68
0.82
0.99
0.79
0.96
1.10
1.41
1.58
1.58
1.23
1.29
1.33
0.63
0.81
1.03
0.65
0.81
0.63
0.80
0.99
1.04
1.32
↑
0.79
0.99
1.30
0.94
1.21
1.52
↑
-
-
↑
↑
↑
0.77
0.99
1.19
0.79
0.99
0.77
0.97
1.15
↑
↑
↑
0.95
1.15
1.37
1.10
1.33
1.54
1.89
2.05
2.10
1.70
1.80
1.92
300
100
50
100
50
100
50
30
300
100
50
300
100
50
100
50
30
100
50
30
100
50
30
1.61
1.68
1.74
3-11µm
3-30µm
5-30µm
5-11µm
-
-
-
-
-
-
8.9
10.5
12.9
10.2
11.8
14.7
14.9
18.2
24.3
12.9
15.1
18.2
-
60
-
-
-
-
-
-
-
-
-
-
72.8
-
-
Table 3. Shown are the number of transits required to de-
tect each molecule to 3.6σ confidence for each of our test
cases. The values in blue indicate the three lowest number
of transits (for each molecule with the exception of CO2)
from the shortest wavelength ranges. A dash indicates that
achieving a 3.6σ was not possible within 100 transits and
therefore, outside the scope of this study.
Figure 9. Compares the improvements of the abundance
constraints for O3, with increasing numbers of transits at
three resolutions for the two key bandpass choices.
this bandpass exceed an order-of-magnitude at 1.41dex
(R=100) and 1.58dex (R=50). If we choose to observe
further into the mid-infrared (5-30µm), including the
Table 4. Shown are the logarithmic constraints on the verti-
cal mixing ratios for each molecule, at 50 transits, for each of
our test cases. The values reported are the half-width of the
posterior distribution, therefore, 1.0 dex indicates the ability
to constrain the abundance to within an order-of-magnitude
higher or lower (e.g., ±1 dex) than the retrieved value. The
values in blue indicate the constraints better than one order-
of-magnitude (for each molecule) from the shortest wave-
length ranges. The ↑ marker denotes an upper limit and dash
represents that the parameter was entirely unconstrained by
the retrieval. The posteriors for all of the R=50 and R=100
test cases will be provided as supplementary material avail-
able at: https://www.dropbox.com/sh/nehnyq6mlkfb13w/
AABuDn-uFqTvAptWVSbGoA04a?dl=0
features at 9.6µm and 14.7µm, we notice only slight re-
ductions in the required observation time and maintain
larger than an order-of-magnitude constraint on both
abundance values. The one noticeable benefit of includ-
ing these longer wavelengths is that less than 20 transits
are required to detect O3 even at a resolution of R=30.
Alternatively, extending the bandpass to 3-11µm to in-
clude the 4.75µm feature, results in a reduction of the
necessary observation time to 10.5 transits for a resolu-
tion of R=100 (12.9 for R=50) and improves our abun-
dance constraints by nearly a factor of four (0.82dex and
0.99dex) for both R=100 and R=50.
It is worth not-
R50R100R3001-5m3-11ming that the 4.75µm feature has significant overlap with
a minor CO2 feature, thereby causing a degeneracy if
observed without additional O3 features. We conclude
that the optimal wavelength coverage for detecting and
constraining O3 is the 3-11µm bandpass choice. Addi-
tionally, this bandpass would only require a resolution
of R=50 to place an order-of-magnitude constraint on
the volume mixing ratio for O3.
3.5. Nitrous Oxide (N2O)
Figure 10. Compares the improvements of the abundance
constraints for N2O, with increasing numbers of transits at
three resolutions for the two key bandpass choices.
Similar to O3, the most prominent features for N2O
extend out into the longer wavelengths. N2O's strongest
spectral features are located at 4.5µm and 7.7µm, each
with an adjacent smaller feature at 3.9µm and 8.6µm,
respectively. Although there is a broad feature at 17µm
it is overlapped by a much larger CO2 feature.
Due to the large degree of degeneracy with the 7.6µm
CH4 feature, isolating the 7.7µm and 8.6µm features in
the 5-11µm bandpass proves to be inadequate. At a
spectral resolution of R=100, the 5-11µm bandpass re-
quires nearly twelve times as many transits (∼94 tran-
sits) as the 3-11µm bandpass would require to obtain
a 3.6σ detection. Extending to 30µm incorporates the
17µm feature and only reduces the observational time to
72.8 transits while the constraints still exceed 1.50dex.
We find that the 3-5µm bandpass is sufficient to detect
N2O at 14.4 transits (R=100) or 23.8 transits (R=50),
making it one of the only two molecular species that can
be constrained in that bandpass within relatively few
transits. However, that region is too narrow to provide
adequate abundance constraints, resulting in a 1.27dex
constraint at 50 transits. If we expand to either the 1-
5µm or the 3-11µm bandpass, our required number of
transits drops to 9.2 and 7.9, respectively. The corre-
sponding constraints at 50 transits are 0.97dex (1-5µm)
11
and 1.02dex (3-11µm), roughly an order-of-magnitude
for both. The 3-11µm region would thus allow for a
3.6σ detection in fewer transits than would be required
for either O3 or H2O within the same regime. Likewise,
the 1-5µm bandpass requires only 39.5% of the obser-
vational time, to detect N2O, as needed to detect H2O
in the same regime. With comparable abundance con-
straints, the 3-11µm bandpass proves to be the optimal
bandpass, requiring only 74% to 86% (R=50 to R=100)
of the observation time as the 1-5µm bandpass choice.
4. DISCUSSION
Unsurprisingly, we find that obtaining the best detec-
tion significance values and abundance constraints for
all of the molecules of interest are achieved by including
the broadest possible wavelength coverage at the high-
est possible spectral resolution. Our primary goal, how-
ever, was to determine how the detection significance
and abundance constraints behave as a function of the
continuous wavelength coverage and spectral resolution.
In the previous section, we elaborate on the effect that
additional wavelength coverage and finer resolution have
on the retrieval results for each molecule individually.
Below we summarize our key findings:
• When considering the contribution of additional
wavelength coverage to the capability of an in-
strument with a continuous infrared bandpass, the
most crucial spectral features (for H2O, CO2, CH4,
O3, N2O, and CO) in transmission spectra do not
extend beyond 11µm.
• Utilizing a spectral resolution greater than R=100
proves to be significantly beneficial to resolving
features at wavelengths shorter than 5µm. Redder
wavelengths offer notably broader features, easily
resolved with a resolution of R=50.
• Degeneracies caused by significantly overlapping
spectral features are most easily resolved by in-
cluding an additional
feature for one or both
molecules.
A brief summary of the crucial molecule-specific take-
aways:
• When attempting to detect and constrain CH4,
the observation of the 3.3µm feature is essential.
If one must opt for lower resolutions (R=100), in-
cluding the 2.3µm feature is incredibly valuable.
While it is certainly ideal to include both features,
it is crucial to acknowledge that the benefit of a
higher resolution (R=300), in the absence of the
2.3µm feature, outweighs the benefit of including
this feature at a lower resolution (R=100).
R50R100R3001-5m3-11m12
Tremblay et al.
• The results of this study validate an alternative
to the observing strategies which are currently
available for H2O, which are limited to the weak
near-infrared features. We find that the more
prominent unobscured spectral feature centered at
6.3µm, combined with the dense cluster of features
from other molecules at 3-5µm, provides a signif-
icant decrease in the necessary observing time to
detect H2O compared to the 1-5µm bandpass.
• CO2, due to its very prominent features, is eas-
ily detectable even only observing the 4.3µm fea-
ture. However,
increasing the wavelength cov-
erage has statistically significant effects on the
abundance constraints regardless of whether the
broader bandpass includes additional CO2 fea-
tures.
• Given a mK = 8.0 source star, we find that utiliz-
ing an instrument with wavelength coverage from
3-11µm, a spectral resolution of only R=50, and a
25m2 collecting area is sufficient to detect ozone
abundances representative of modern-day Earth
with less than 13 transits and to constrain the
abundance value within an order-of-magnitude at
50 transits.
• All of the N2O features from 1-30µm are partially
or completely degenerate with features from CO2
or CH4. However, under the detector set-up eval-
uated in this study, observing only the features in
the 3-5µm regime allows for a statistically signifi-
cant detection in less than 15 transits. The 1-5µm
and the 3-11µm bandpasses offer comparable im-
provements to both the detection significance and
abundance constraints.
The ability to detect and constrain the five biolog-
ically relevant molecular species in the atmosphere of
an Earth-analog should be the benchmark by which the
next-generation of exoplanet observatories are designed.
The detection of significant amounts of CH4 in tandem
with O3 and/or N2O would constitute a promising in-
dicator of biogenic origins. However, as it has been la-
boriously stated, CH4 is the most challenging molecule
to detect in this study and will prove to be the limit-
ing variable in the development of a detector for these
purposes. Unlike the other molecules considered here,
CH4 does not possess broad unobscured features in the
mid-infrared and therefore does not have the advantage
of relaxed spectral resolution requirements. It is impor-
tant to acknowledge, however; that this is not an argu-
ment against observing mid-infrared wavelengths, which
have been shown to be immensely valuable in the detec-
tion and constraint of all of the other molecular species
in this study. Rather, our intention is to highlight that
the choice of wavelength coverage for a next-generation
near-to-mid-infrared spectrometer, particularly on the
blue end of the bandpass, must be influenced by one or
both of the near-infrared CH4 features. To that end,
we conclude that a near-to-mid-infrared spectrometer,
aboard a next-generation space-based telescope (with a
collecting area of 25m2), boasting a bandpass including
features between ∼2-11µm and a resolution of R(cid:39)50-
300 would prove capable of detecting and constraining
all of the key bio-indicator gasses in Earth's atmosphere.
While the 2-11µm bandpass was not explicitly probed
in this study, our retrieval results indicate that it would
provide 3.6σ detections in less transits and narrower
abundance constraints than either the 1-5µm or 3-11µm
bandpasses studies here.
5. CONCLUSION
Our work has been motivated by the desire within the
exoplanet community to characterize the atmospheres
of temperate terrestrial planets beyond our Solar Sys-
tem. The detection of molecular biosignatures on these
small rocky planets looms just beyond the capabilities
of current space-based telescopes. A future facility with
broad continuous wavelength coverage in the near and
mid-infrared combined with detectors capable of driv-
ing observational noise down to the astrophysical noise
floor will have what it takes to comprehensively probe
the climate and composition of terrestrial exoplanet at-
mospheres.
The 2020 Decadal Survey will consider four flagship
mission concepts including one of particular interest to
this study, the Origins Space Telescope. The Origins
concept includes a near-to-mid-infrared spectrograph
boasting an impressive 2.8-20µm continuous bandpass
ranging from R(cid:39)50-300 with detectors designed to re-
duce instrument noise to approximately 5ppm (OST-
STDT5). Although it excludes the 2.3µm feature for
CH4, the design of this instrument strongly aligns with
the findings in this work for the optimal observational
setup for both detecting the presence and constraining
the abundances of Earth-like biosignatures.
We look to expand upon this work by testing a sub-
set of this analysis to a broader range of atmospheric
compositions in order to determine the possibility of
distinguishing several plausible atmospheric composi-
tions on known terrestrial planets and those soon-to-be-
5
https://asd.gsfc.nasa.gov/firs/docs/
OriginsVolume1MissionConceptStudyReport.pdf
discovered by TESS. Additionally, we acknowledge the
benefit of thermal emission spectroscopy for character-
izing the thermal structure of exoplanetary atmospheres
and providing additional information on the molecular
abundances. Therefore, we aim to explore the effect
of these observational parameters on a similar grid of
synthetic thermal emission spectra. To this end, we an-
ticipate that the synthesis of transmission and thermal
emission as well as reflected light spectroscopy to be
essential to the comprehensive understanding of these
exoplanetary environments.
Software:
PySynPhot (Lim et al. 2015), CHIMERA
(Batalha & Line 2017; Line et al. 2013; Greene et al. 2016;
Line & Parmentier 2016)), PyMultiNest (Buchner et al.
2014)
ACKNOWLEDGEMENTS
13
We would like to thank Tom Greene for providing
us with his noise model. M.R. Line and L. Tremblay
acknowledge the NASA XRP award NNX17AB56G for
supporting this work. This work benefited from the 2018
Exoplanet Summer Program in the Other Worlds Lab-
oratory (OWL) at the University of California, Santa
Cruz, a program funded by the Heising-Simons Foun-
dation. The authors also acknowledge NASA, which
through the Origins Space Telescope mission concept
study supported travel to the Other Worlds Laboratory
Summer Program to facilitate this work.
Part of the research was carried out at the Jet Propul-
sion Laboratory, California Institute of Technology, un-
der contract with the National Aeronautics and Space
Administration.
REFERENCES
Barclay, T., Pepper, J., & Quintana, E. V. 2018, ApJS,
Gillon, M., Triaud, A. H. M. J., Demory, B.-O., et al. 2017,
239, 2
Nature, 542, 456
Barstow, J. K., Aigrain, S., Irwin, P. G. J., et al. 2014,
Glass, J. D., & Blair, W. D. 2015, IEEE Transactions on
ApJ, 786, 154
Aerospace Electronic Systems, 51, 1927
Barstow, J. K., & Irwin, P. G. J. 2016, MNRAS, 461, L92
Glass, J. D., Blair, W. D., & Lanterman, A. D. 2015, IEEE
Batalha, N. E., Lewis, N. K., Line, M. R., Valenti, J., &
Transactions on Signal Processing, 63, 6673
Stevenson, K. 2018, ApJL, 856, L34
Gordon, I., Rothman, L., Hill, C., et al. 2017, Journal of
Batalha, N. E., & Line, M. R. 2017, AJ, 153, 151
Quantitative Spectroscopy and Radiative Transfer, 203, 3
Beichman, C., Benneke, B., Knutson, H., et al. 2014, PASP,
Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ,
126, 1134
817, 17
Benneke, B., & Seager, S. 2012, ApJ, 753, 100
Grimm, S. L., Demory, B.-O., Gillon, M., et al. 2018, A&A,
-- . 2013, ApJ, 778, 153
613, A68
Buchner, J., Georgakakis, A., Nandra, K., et al. 2014,
Hardegree-Ullman, K. K., Cushing, M. C., Muirhead, P. S.,
A&A, 564, A125
& Christiansen, J. L. 2019, AJ, 158, 75
Catling, D., & Kasting, J. 2007, Planetary Atmospheres
Harman, C. E., Felton, R., Hu, R., et al. 2018, ApJ, 866, 56
and Life, ed. W. Sullivan & J. Baross (United Kingdom:
Cambridge University Press), 91 -- 116
Harman, C. E., Schwieterman, E. W., Schottelkotte, J. C.,
& Kasting, J. F. 2015, ApJ, 812, 137
Chapman, J. W., Zellem, R. T., Line, M. R., et al. 2017,
Husser, T. O., Wende-von Berg, S., Dreizler, S., et al. 2013,
PASP, 129, 104402
A&A, 553, A6
Cockell, C. S., Kaltenegger, L., & Raven, J. A. 2009,
Kasting, J. F. 2009, Geochimica et Cosmochimica Acta
Astrobiology, 9, 623
Supplement, 73, A625
Cowan, N. B., Greene, T., Angerhausen, D., et al. 2015,
Kempton, E. M. R., Bean, J. L., Louie, D. R., et al. 2018,
PASP, 127, 311
PASP, 130, 114401
Deming, D., Seager, S., Winn, J., et al. 2009, PASP, 121,
Kochanov, R., Gordon, I., Rothman, L., et al. 2016, Journal
952
Domagal-Goldman, S. D., Segura, A., Claire, M. W.,
Robinson, T. D., & Meadows, V. S. 2014, ApJ, 792, 90
of Quantitative Spectroscopy and Radiative Transfer,
177, 15 , xVIIIth Symposium on High Resolution
Molecular Spectroscopy (HighRus-2015), Tomsk, Russia
Gillon, M., Jehin, E., Lederer, S. M., et al. 2016, Nature,
Krissansen-Totton, J., Bergsman, D. S., & Catling, D. C.
533, 221
2016, Astrobiology, 16, 39
14
Tremblay et al.
Krissansen-Totton, J., Garland, R., Irwin, P., & Catling,
Sagan, C., Thompson, W. R., Carlson, R., Gurnett, D., &
D. C. 2018, AJ, 156, 114
Hord, C. 1993, Nature, 365, 715
L´eger, A. 2000, Advances in Space Research, 25, 2209
L´eger, A., Pirre, M., & Marceau, F. J. 1993, A&A, 277, 309
Lim, P., Diaz, R., & Laidler, V. 2015, Astrophysics Source
Code Library
Line, M. R., & Parmentier, V. 2016, ApJ, 820, 78
Line, M. R., Zhang, X., Vasisht, G., et al. 2012, ApJ, 749,
93
Line, M. R., Wolf, A. S., Zhang, X., et al. 2013, ApJ, 775,
137
Lustig-Yaeger, J., Meadows, V. S., & Lincowski, A. P. 2019,
AJ, 158, 27
Schwieterman, E. W., Robinson, T. D., Meadows, V. S.,
Misra, A., & Domagal-Goldman, S. 2015, ApJ, 810, 57
Schwieterman, E. W., Meadows, V. S., Domagal-Goldman,
S. D., et al. 2016, ApJL, 819, L13
Seager, S. 2014, Proceedings of the National Academy of
Science, 111, 12634
Seager, S., & Bains, W. 2015, Science Advances, 1,
e1500047
Seager, S., & Deming, D. 2010, ARA&A, 48, 631
Segura, A., Krelove, K., Kasting, J. F., et al. 2003,
Astrobiology, 3, 689
Stevenson, K. B., Lewis, N. K., Bean, J. L., et al. 2016,
Matsuo, T., Greene, T., Roellig, T. L., et al. 2018, in Space
PASP, 128, 094401
Telescopes and Instrumentation 2018: Optical, Infrared,
and Millimeter Wave, Vol. 10698, International Society
for Optics and Photonics (SPIE), 1218 -- 1229
Suissa, G., Chen, J., & Kipping, D. 2018, MNRAS, 476,
2613
Suissa, G., & Kipping, D. 2018, Research Notes of the
Meadows, V. S. 2017, Astrobiology, 17, 1022
American Astronomical Society, 2, 31
Meadows, V. S., Reinhard, C. T., Arney, G. N., et al. 2018,
Tian, F., France, K., Linsky, J. L., Mauas, P. J. D., &
Astrobiology, 18, 630
Morley, C. V., Kreidberg, L., Rustamkulov, Z., Robinson,
T., & Fortney, J. J. 2017, ApJ, 850, 121
Quirrenbach, A., Amado, P. J., Ribas, I., et al. 2018, in
Society of Photo-Optical Instrumentation Engineers
(SPIE) Conference Series, Vol. 10702, Proc. SPIE,
107020W
Robinson, T. D., Fortney, J. J., & Hubbard, W. B. 2017,
ApJ, 850, 128
Rocchetto, M., Waldmann, I. P., Venot, O., Lagage, P. O.,
& Tinetti, G. 2016, ApJ, 833, 120
Vieytes, M. C. 2014, Earth and Planetary Science
Letters, 385, 22
Trotta, R. 2008, Contemporary Physics, 49, 71
Turbet, M., Bolmont, E., Leconte, J., et al. 2018, A&A,
612, A86
Unterborn, C. T., & Panero, W. R. 2017, ApJ, 845, 61
Wolf, E. T. 2017, ApJL, 839, L1
Wunderlich, F., Godolt, M., Grenfell, J. L., et al. 2019,
A&A, 624, A49
Zechmeister, M., Dreizler, S., Ribas, I., et al. 2019, A&A,
627, A49
Appendices
15
Figure 11. Corner plot for the 3-11µm at R=100 test case. Shows the posterior probability distributions for all of the
parameters in our model. In the dropbox link (presented earlier) we provide all of the corner plots for the R=50 and R=100
scenarios.
Tiso = 281.11+22.85−21.250.751.001.251.50×Rp×Rp = 1.00+0.00−0.004.53.01.50.01.5log(CTP)log(CTP) = −0.42+0.58−0.5810203040Bkg MMWBkg MMW = 29.02+3.48−3.0810.07.55.02.50.0log(H2O)log(H2O) = −5.23+1.16−1.1310.07.55.02.50.0log(CH4)log(CH4) = −6.20+0.97−1.0010.07.55.02.50.0log(CO2)log(CO2) = −3.16+1.10−1.0510.07.55.02.50.0log(O3)log(O3) = −6.30+0.86−0.7810.07.55.02.50.0log(N2O)log(N2O) = −6.04+1.05−0.99200400600800Tiso10.07.55.02.50.0log(CO)0.751.001.251.50×Rp4.53.01.50.01.5log(CTP)10203040Bkg MMW10.07.55.02.50.0log(H2O)10.07.55.02.50.0log(CH4)10.07.55.02.50.0log(CO2)10.07.55.02.50.0log(O3)10.07.55.02.50.0log(N2O)10.07.55.02.50.0log(CO)log(CO) = −8.77+2.08−2.18 |
1912.11149 | 1 | 1912 | 2019-12-24T00:03:03 | High-Resolution Transmission Spectra of Earth through Geological Time | [
"astro-ph.EP",
"astro-ph.IM",
"astro-ph.SR"
] | The next generation of ground- and space-based telescopes will be able to observe rocky Earth-like planets in the near future, transiting their host star. We explore how the transmission spectrum of Earth changed through its geological history. These transmission spectra provide a template for how to characterize an Earth-like exoplanet - from a young prebiotic world to a modern Earth. They also allow us to explore at what point in its evolution a distant observer could identify life on our Pale Blue Dot and other worlds like it. We chose atmosphere models representative of five geological epochs of Earth's history, corresponding to a prebiotic high CO2-world 3.9 billion years ago (Ga), an anoxic world around 3.5 Ga, and 3 epochs through the rise of oxygen from 0.2 percent to present atmospheric levels of 21 percent.
Our transmission spectra show atmospheric spectral features, which would show a remote observer that Earth had a biosphere since about 2 billion years ago. These high-resolution transmission spectral database of Earth through geological time from the VIS to the IR is available online and can be used as a tool to optimize our observation strategy, train retrieval methods, and interpret upcoming observations with JWST, the Extremely Large Telescopes and future mission concepts like Origins, HabEx, and LUOVIR. | astro-ph.EP | astro-ph | HIGH-RESOLUTION TRANSMISSION SPECTRA OF EARTH THROUGH
GEOLOGICAL TIME
Lisa Kaltenegger1, 2, Zifan Lin1,2 & Jack Madden1,2
1Cornell University, Astronomy and Space Sciences Building, Ithaca, NY 14850, USA
2Carl Sagan Institute, Space Science Building 311, Ithaca, NY 14850, USA
ABSTRACT
The next generation of ground- and space-based telescopes will be able to observe
rocky Earth-like planets in the near future, transiting their host star. We explore how
the transmission spectrum of Earth changed through its geological history. These
transmission spectra provide a template for how to characterize an Earth-like
exoplanet -- from a young prebiotic world to a modern Earth. They also allow us to
explore at what point in its evolution a distant observer could identify life on our Pale
Blue Dots and other worlds like it. We chose atmosphere models representative of
five geological epochs of Earth's history, corresponding to a prebiotic high CO2-
world 3.9 billion years ago (Ga), an anoxic world around 3.5 Ga, and 3 epochs
through the rise of oxygen from 0.2% to present atmospheric levels of 21%.
Our transmission spectra show atmospheric spectral features, which would show a
remote observer that Earth had a biosphere since about 2 billion years ago. These
high-resolution transmission spectral database of Earth through geological time from
the VIS to the IR is available online and can be used as a tool to optimize our
observation strategy, train retrieval methods, and interpret upcoming observations
with JWST, the Extremely Large Telescopes and future mission concepts like
Origins, HabEx, and LUOVIR.
INTRODUCTION
Among the more than 4000 discovered
exoplanets to date are dozens of Earth-size
planets (see, e.g., Udry et al. 2007;
Borucki et al. 2011, 2013; Kaltenegger &
Sasselov 2011; Batalha et al. 2013;
Kaltenegger et al. 2013; Quintana et al.
2014; Torres et al. 2015), including
2017, John et al 2018, Berger et al 2018).
several with similar irradiation to Earth
(see e.g., Kane et al 2016, Kaltenegger
The space-based James Webb Space
telescope (JWST) is scheduled to launch
in early 2021 and several ground-based
Extremely Large Telescopes (ELTs) are
construction or
currently under
in
the Giant Magellan
like
planning,
Telescope
(GMT),
the Thirty Meter
Telescope (TMT), and
the Extremely
(ELT), which are
Large Telescope
designed to be able to undertake the first
measurements of
the atmospheres of
Earth-sized planets (see e.g. Kaltenegger
& Traub 2009, Kaltenegger et al 2010,
Garcia-Munoz et al 2012, Hedelt et al.
2013, Snellen et al 2013, Rodler & Lopez-
Morales 2014, Betremieux & Kaltenegger
2014, Misra et al 2014, Stevenson et al
2016, Barstow et al 2016, Kaltenegger et
al 2019, Zin & Kaltenegger 2019). Several
future mission concepts
like Origins
(Battersby et al 2018), Habex (Mennesson
et al. 2016) and LUOVIR (LuvoirTeam et
al 2018) are currently being designed to be
able
atmospheric
composition of Earth-sized planets.
explore
Earth's atmosphere has undergone a
substantial evolution since formation (see
e.g. Walker 1977, Zahnle et al 2007,
Lyones et al 2014). Previous work by one
the
to
of
the
and
representative
of the authors modeled Earth's reflection
and emission spectra through geological
time,
exoplanet
observations seen as directly imaged Pale
Blue Dots (Kaltenegger et al. 2007). A
second paper including one of the authors
investigated how
reflection and
emission spectra of Earth through its
geological history from anoxic to modern
Earth-like planets changes if they are
orbiting different Sun-like host stars from
F0V to M8V spectral type (Rugheimer &
Kaltenegger 2018). The surface UV
environment for Earth through geological
time for our Sun and around different Sun-
like host stars shows comparable UV
surface environments for such planets as
discussed in Rugheimer & Kaltenegger
(2018)
&
Kaltenegger (2017, 2019).
O'Malley-James
While emission and reflection spectra
for models of Earth through geological
time exist (e.g. Kaltenegger et al 2007,
Rugheimer & Kaltenegger
2018),
transmission spectra have been focused on
modern Earth so far (see e.g., Ehrenreich
et al. 2006, Kaltenegger & Traub 2009,
Palle et al. 2009, Vidal-Madjar et al. 2010,
Rauer et al. 2011, Garcıa Munoz et al.
2012, Hedelt et al. 2013, Betremieux &
Kaltenegger 2013, 2014, Misra et al.
2014).
Here we model the high-resolution
transmission spectra for five geological
epochs
in Earthʼs history (Table 1):
representative of a high-CO2 prebiotic
world as epoch 1 around 3.9 Ga, an
Anoxic world as epoch 2 around 3.5 Ga
and 3 epochs during the rise of oxygen,
corresponding to the timeframe of the rise
of oxygen in Earth's atmosphere between
about 2.4 Ga to today (see review by
Lyons et al 2014). We modeled epoch 3
after the Grand Oxygenation Event, with
1% PAL (present atmospheric levels) of
O2, epoch 4 after the Neoproterozoic
Our
Oxygenation Event, with 10% PAL O2.
Epoch 5
represents modern Earth
atmosphere with 21% O2. We use a solar
evolution model (Claire et al. 2012) to
establish the incident Solar Flux through
Earth's geological evolution.
database
high-resolution
of
transmission spectra from the visible to the
Infrared (0.4 µm to 20 µm) for Earth
through geological time, which is freely
online
available
(www.carlsaganinstitute.org/data),
is
a
tool to enable effective observations and
first interpretation of atmospheric spectra
of Earth-like planets, using our planet's
evolution as a template. Our models and
transit spectra include climate indicators
like H2O and CO2 as well as biosignatures
like the combination of O2 or O3 in
combination with a reduced gas like CH4
(see discussion on biosignatures e.g. in
Kaltenegger 2017). The term biosignatures
is used here to mean remotely detectable
atmospheric gases that are produced by
life and are not readily mimicked by
abiotic processes, e.g. the CH4 + O2
(Lederberg 1965, Lovelock 1965) or CH4
+ N2O (Lippincott et al. 1967) pairs.
In section 2 we describe our model. In
section 3 we present the transmission
spectra for 5 epochs in Earth's geological
history and discuss the absorption features
of climate indicators and biosignatures in
low and high-resolution from the visible to
infrared wavelength. Section 4 discusses
and summarizes our results.
METHODS
We used EXO-Prime (for details see
Kaltenegger et al 2007, Kaltenegger &
Sasselov 2010 and Madden & Kaltenegger
2019) to simulate Earth's atmosphere and
transmission spectra for 5 geological
epochs. EXO-Prime is a coupled 1D
iterative climate-photochemistry code (see
e.g. Kasting & Ackerman 1986, Pavlov &
Kasting 2002, Segura et al. 2005, 2007,
Haqq-Misra et al. 2008, Arney et al. 2016,
Madden & Kaltenegger 2019), with a line
by line Radiative Transfer code (e.g.
Traub & Stier 1976; Kaltenegger & Traub
2009) for rocky exoplanets, which was
originally developed for Earth. EXO-
Prime has been validated for visible to
infrared wavelengths by comparison to
Earth observed as an exoplanet from
different missions like EPOXI, the Mars
Global Surveyor, Shuttle data, and
multiple
observations
(Kaltenegger et al. 2007; Kaltenegger &
Traub 2009; Rugheimer et al. 2013).
earthshine
calculate
high-resolution
transmission spectra at a resolution of 0.01
cm-1 using opacities
the 2016
HITRAN database (Gordon et al. 2017)
for O2, O3, H2O, CO2, CH4, N2O, CH3Cl,
SO2, H2S, H2O2, OH, HO2, HOCl, H2CO,
HCl, ClO, NO2, NO, HNO3, and CO. For
CFCl3 (Sharpe et al. 2004) and N2O5
(Wagner & Birk 2003) we use cross-
sections. We include CO2 line mixing (see
also Niro et al. 2005a, 2005b). For CO2,
H2O, and N2, we use measured continua
data instead of line-by-line calculations in
the far wings (see Traub & Jucks 2002).
from
We
the
Earth's atmosphere cannot be probed in
primary transit below 12 km, because
refraction from the deeper atmospheric
regions deflects light away from a distant
observer in an Earth -- Sun geometry (see
e.g. Garcıa Munoz et al. 2012, Betremieux
& Kaltenegger 2014, Misra et al. 2014).
Our transmission spectra show the cutoff
due to refraction at 12km for all epochs.
Clouds do not significantly affect the
strengths of the spectral features in Earth's
transmission because most clouds on Earth
are located at altitudes below 12km.
through geological
We base the transmission spectra for
times on
Earth
atmosphere models by Kaltenegger et al.
(2007) and Rugheimer et al. (2018): The
models for each epoch are discussed in
detail in those two papers and summarized
in Table 1 and Figure 1. The changing
solar constant accounts for the lower solar
incident flux at earlier times in Earth's
history, following the prediction of a 30%
reduction in solar flux for a young Earth at
4.6 Ga (Claire et al 2012). All epochs
assume a 1 bar surface pressure, consistent
with geological evidence
for paleo-
pressures close to modern values (e.g.
Somet al. 2012; Marty et al. 2013).
Table 1: Chemical mixing ratios for major atmospheric gases in our model atmospheres, for 5
epochs through Earth's geological history from prebiotic to anoxic atmospheres representative of
3.9 Ga and 3.5 Ga in Earth's history to 3 models which capture the rise of oxygen from 0.01PAL
O2 to modern Earth with 21% O2.
CO2
Epoch
5
4
3
2
1
3.65E-04
1.00E-02
1.00E-02
1.00E-02
1.00E-01
Solar
Constant
1.00
0.95
0.87
0.77
0.75
Time
Period
now
0.5 - 0.8
1.0 - 2.0
3.5
3.9
Epoch 1 is a CO2-rich atmosphere of a
prebiotic world, representative of Early
Earth around 3.9 Ga. Epoch 2 is an
Archaen world, representative of a young
Earth around 3.5 Ga. Epoch 3 corresponds
to a Paleo- and Meso-proterozoic Earth
CH4
O2
O3
N2O
1.65E-06
4.15E-04
1.65E-03
1.65E-03
1.65E-06
2.10E-01
2.10E-02
2.10E-03
1.00E-13
1.00E-13
3.00E-08
2.02E-08
7.38E-09
2.55E-19
2.55E-19
3.00E-07
9.15E-08
8.37E-09
0
0
(about 2 to 1 Ga), when oxygen started to
rise in Earthʼs atmosphere. We use 0.21%
O2 (1% PAL) for this model. Epoch 4
corresponds
the proliferation of
multicellular life on Neoproterozoic Earth
(about 0.8 to 0.5 Ga) when the oxygen
to
concentration had risen to 10% PAL (2.1%
O2). Epoch 5 corresponds to modern Earth
with 21% O2. Note that the time of the
oxygen rise has recently been moved to a
later stage in Earth's evolution (see e.g.
review in Lyones et al 2014), which is
reflected in the time ranges gives in Table
1 for epoch 3 and epoch 4, instead of
geological times given in our earlier paper
(Kaltenegger et al. 2007), which based O2
concentrations on work by Holland et al.
(2006).
Oxic Earth
1PAL O2
0.1PAL O2
0.01PAL O2
Anoxic Earth
3.5Ga
3.9Ga
70
60
50
40
30
20
10
)
m
k
(
t
h
g
i
e
H
0
100
200
300
10 − 6
log10(H2O)
10 − 12 10 − 10 10 − 8 10 − 6 10 − 4 10 − 2
150
250
Tem perature (K)
Fig.1: Temperature and mixing ratios for major atmospheric gases in our model atmospheres,
representative of 5 epochs through Earth's geological evolution from an CO2-rich prebiotic
atmosphere for Earth around 3.9Ga to an anoxic atmosphere around 3.5Ga and 3 models, which
capture the rise of oxygen from 0.01PAL O2 to 1PAL (21% O2) on modern Earth. The mixing
ratios shown from left to right are for H2O, CH4, O3 and N2O (see also Table 1).
10 − 5
log10(CH4 )
log10(N2O)
log10(O3)
10 − 10
10 − 12
10 − 10
10 − 8
10 − 4
10 − 3
− 9
10
10 − 8
The transmission spectra in figures 2
are smoothed with a triangular kernel to a
resolving power of 700 for clarity. We did
not add noise to the spectra to provide
theoretical
several
instruments, which all have different
instrument-specific noise profiles which
can easily be added to our model to
provide realistic observation simulations.
spectra
input
for
RESULTS
All
transmission spectra are available
online for a minimum resolution of λ/Δλ >
100,000 for the full wavelength range
from 0.4µm to 20µm (0.01cm-1 steps).
Figure 2 shows the transmission spectra
for a resolution of λ/Δλ = 700 for clarity,
with the most prominent spectral features
identified. The five atmosphere models are
representative of Earth through geological
time sorted from modern Earth on top to
Early Earth on the bottom:
Modern Earth's transmission spectrum is
shown as the top row (Epoch 5, 21% O2),
followed by the transmission spectrum for
Neoproterozoic Earth (Epoch 4, 0.5 to
0.8Ga, O2 = 2.1 x 10-2 (10% PAL)),
followed by the transmission spectrum for
a Paleo- and Meso-proterozoic Earth
(Epoch 3, 1 to 2Ga, O2 = 2.1 x 10-3 (1%
PAL)), an anoxic Earth (Epoch 2, 3.5 Ga)
and a prebiotic Earth (epoch 1, 3.9 Ga),
which is shown in the bottom row.
Throughout the atmospheric evolution
of our Earth model, different absorption
features dominate Earth's spectrum with
CH4 and CO2 being dominant in Early
Earth models, where
they are more
abundant and O2 and O3
features
increasing with O2 abundance from a
Paleo- and Meso-proterozoic Earth in
epoch 3 to modern Earth in epoch 5
(Table 1).
Note that several features overlap at the
resolution of λ/Δλ = 700, which is shown
in Fig. 2 for clarity and are not specifically
labeled. In
the high-resolution online
transmission spectra individual spectral
lines can be easily discerned for these
molecules, as shown in Fig. 3 for O2 and
O3 for biotic atmospheres (epoch 5 to
epoch
of
λ/Δλ = 100,000 as proposed for several
instruments on the ELTs.
resolution
for
3),
a
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
0
Modern Earth: 21% Oxygen
O3
O2
O2
H2O
g'
r'
i'
Z
Y
J
H
NIRSpec/NIRISS
Earth: Rise of Oxygen (2.1% O2)
85.50
85.25
85.00
84.75
CO2
H2O
84.50
O3
O2
O2
H2O
H2O
CO2
H2O
CO2
CH4
H2O
g'
r'
i'
Z
Y
J
H
NIRSpec/NIRISS
Earth: Rise of Oxygen (0.2% O2)
H2O
H2O
CO2
CO2
CH4
H2O
H2O
O2
g'
r'
i'
Z
Y
J
H
NIRSpec/NIRISS
Anoxic Earth: 3.5 Billion years ago
H2O
CH4
H2O
H2O
g'
r'
i'
Z
Y
J
Anoxic Earth: 3.9 Billion years ago
CO2
CH4
H2O
CO2
H
NIRSpec/NIRISS
CH4
CO2
CO2
g'
r'
i'
Z
Y
J
H
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
Modern Earth: 21% Oxygen
H2O
CO2
CH4
K
CH4
L
NIRSpec/NIRISS
Earth: Rise of Oxygen (2.1% O2)
CO2
CH4
K
H2O
CH4
L
NIRSpec/NIRISS
Earth: Rise of Oxygen (0.2% O2)
H2O
CH4
40
CO2
CH4
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
K
L
NIRSpec/NIRISS
Anoxic Earth: 3.5 Billion years ago
H2O
CH4
CO2
CH4
K
L
NIRSpec/NIRISS
Anoxic Earth: 3.9 Billion years ago
H2O
CO2
CH4
K
CH4
L
CO2
CO2
CO2
CO2
CO2
CO2
M
CO2
M
CO2
M
CO2
M
CO2
M
)
m
p
p
(
*
R
/
P
R
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
p
p
(
*
R
/
P
R
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
p
p
(
*
R
/
P
R
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
p
p
(
*
R
/
P
R
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
)
m
p
p
(
*
R
/
P
R
e
v
i
t
c
e
ff
E
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.25
)
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
i
t
c
e
ff
E
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
0
60
50
40
30
20
10
0
)
m
p
p
(
*
R
/
P
R
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
k
(
t
h
g
i
e
h
e
v
i
t
c
e
ff
E
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
Modern Earth: 21% Oxygen
CO2
O3
N2O
CO2
H2O
CH4
60
50
40
30
)
m
k
(
t
h
g
i
e
h
e
v
)
m
p
p
(
*
R
/
P
R
N2O
N2O
N
MIRI
Earth: Rise of Oxygen (2.1% O2)
CO2
CH4
CO2
H2O
O3
CO2
N
MIRI
Earth: Rise of Oxygen (0.2% O2)
CO2
CH4
CO2
H2O
O3
CO2
N
MIRI
Anoxic Earth: 3.5 Billion years ago
CO2
CH4
CO2
H2O
CO2
N
MIRI
Anoxic Earth: 3.9 Billion years ago
CO2
CO2
CH4
H2O
CO2
N
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
85.50
85.25
85.00
84.75
84.50
84.25
84.00
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
)
m
p
p
(
*
R
/
P
R
H2O
Q
H2O
Q
H2O
Q
H2O
Q
H2O
Q
NIRSpec/NIRISS
0
NIRSpec/NIRISS
MIRI
9
8
7
6
5
20
19
18
17
16
15
14
12
11
10
13
4.0
5.0
4.5
3.5
3.0
2.5
2.0
Wavelength ( μm )
Wavelength ( μm )
Wavelength ( μm )
0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0
Fig. 2: Model Spectra for Earth through geological time from 0.4 to 20 µm shown at a resolution
of λ/Δλ = 700 for 5 epochs through Earth's geological time from an anoxic atmosphere 3.9 Ga to
an anoxic atmosphere with less CO2 and CH4 around 3.5 Ga and 3 models which capture the rise
of oxygen from 0.01 PAL O2 to 1PAL (21% O2) on modern Earth, which started around 2.4
billion years ago (Ga).
Absorption
the visible
wavelength range (left panel, 0.4 to 2
µm): Modern Earth (epoch 5, top row)
shows a strong absorption feature for O2 at
0.76 µm, with a weaker feature at 0.69 µm.
O3
from
approximately 0.45 to 0.74 µm. These
features decrease with decreasing oxygen
composition for younger Earth models and
disappear for anoxic Earth atmosphere
models (epoch 1 and epoch 2). CH4 in
modern Earth's atmosphere shows no
significant visible absorption features in
transmission in the visible in Fig. 2, but at
higher abundance, it shows absorption
features at 1.7, and also at 0.88 and 1.04
a broad
features
feature,
shows
in
µm in early Earth models. H2O shows
absorption features at a wide range of
wavelength in the visible at 0.73, 0.82,
0.95, and 1.14 µm. CO2 does not show
visible features at present abundance, but
in a high-CO2 atmosphere of 10% CO2, as
in early Earth evolution stages, the weak
1.06, 1.3, 1.6 and 2µm features can be
seen in Figure 2.
in
features
Absorption
the NIR
wavelength range (middle panel from 2
to 5 µm): Neither O2 nor O3 show
absorption features in the NIR. CH4 shows
absorption features at 2.4 and 3.3 µm,
which increase with CH4 abundance in
Earth's atmosphere for earlier geological
epochs. Several CO2 features can be
identified in Fig. 2 with increasing CO2
abundance for younger Earth models. H2O
abundance and absorption feature strength
increase
surface
temperature and consequence evaporation
rate for epoch 3 and epoch 4.
increasing
with
an
and
therefore
Absorption in the IR wavelength range
(right panel from 5 to 20 mm): At 9.6 µm
the strength of the absorption feature of O3
decreases with decreasing O3 abundance
for younger oxic Earths. It is a saturated
feature
excellent
qualitative but poor quantitative indicator
for the existence of O2. A smaller O3
feature can be seen at 9µm also decreasing
in strength with decreasing O3 abundance.
For anoxic atmospheres the O3 feature is
not visible. At 7.6 µm the absorption
feature of CH4 becomes stronger with
increasing CH4 abundance for younger
Earth models. The main CO2 feature at
15µm as well as several smaller CO2
features at 10.4 and 9.4 µm increase with
increasing CO2 abundance for earlier Earth
models. H2O features can be seen at 7 and
20 µm. N2O has spectral features at 16.89
µm, which can be seen in epoch 3 and 4 in
the wing of the 15 µm CO2 feature, with
smaller features at 7.75, 8.52, and 10.65
µm, which overlap with other spectral
features at the resolution shown in Fig. 2.
1 PAL O2
0.1 PAL O2
0.01 PAL O2
1 PAL O2
0.1 PAL O2
0.01 PAL O2
Fig. 3: High-resolution (λ/Δλ > 100,000) for the 0.76 µm O2 and 9.6 µm O3 feature for the rise of
oxygen from 0.01 to the present atmospheric level (PAL) of oxygen, which is 21% in Earth's
modern atmosphere.
spectra
Note that in the online high-resolution
transmission
some absorption
features, which are not apparent in Figure 2
at a resolution of λ/Δλ = 700, can be
identified. As an example we show the
change in both the O2 feature at 0.76 µm
and the O3 feature at 9.6 µm through
geological time in Figure 3 for a minimum
resolution of λ/Δλ = 100,000 for the whole
wavelength range shown.
Discussion and Conclusion
high-resolution
We
transmission
database
of
atmospheric models representative of Earth
through its geological history from the VIS
to the IR (0.4 to 20µm) with a minimum
resolution of 100,000. These transmission
spectra provide a template of how to
remotely characterize Earth-like exoplanets
with upcoming ground- and space-based
telescopes and explore at what point in its
evolution a distant observer could identify
life on our own planet and others like it.
a
spectral
generated
We chose atmospheres representative of
five geological epochs of Earth's history,
corresponding to a prebiotic high CO2-
world and an anoxic world at 3.9 and 3.5
billion years ago, as well as 3 epochs
through the rise of O2 from 1% of to
present atmospheric levels, which started
2.4 billion years ago on Earth. Throughout
the atmospheric evolution of our Earth,
different absorption
features dominate
Earth's transmission spectrum (shown in
Fig. 2 at a resolution of λ/Δλ = 700) with
CH4 and CO2 being dominant in Early
Earth models, where
they are more
abundant. O2 and O3 spectral features
become stronger with increasing abundance
during the rise of oxygen (Epoch 3 to 5).
Analyzing the emergent spectrum of
Earth, taken by the Galileo probe, Sagan et
al. (1993) concluded that the large amount
of O2 in the presence of CH4 is strongly
suggestive of biology, as Lovelock (1965)
is
in
species
and Lederberg
(1965) had suggested
earlier. On short timescales, the two species
react
to produce CO2 and H2O and
therefore, must be constantly replenished to
maintain detectable concentrations (see e.g.
review Kaltenegger 2017). It
their
quantities and detection along with other
atmospheric
the planetary
context that solidify a biological origin (as
discussed in detail in several recent reviews
e.g. Kasting et al. 2014, Kaltenegger 2017,
Swieterman
the
combination of these gases can be detected
in several different wavelengths, dependent
on their abundance and Earth's geological
evolution, as shown in Fig. 2. The strongest
features in the visible are O2 (0.76 µm) and
O3 (0.6 µm), in the NIR CH4 (2.4µm), and
in the thermal IR O3 (9.6µm) and CH4 (7.6
µm). High-resolution spectral features for a
minimum resolution of λ/Δλ = 100,000 are
shown in Fig. 3 for O2 and O3.
2018). On Earth,
strategy,
(λ/Δλ >100,000)
The high resolution transmission spectra
database
is available
online www.carlsaganinstitute.org/data and
can be used as a tool to optimize our
observation
retrieval
methods, as well as interpret upcoming
observations with JWST as well as ground-
based Extremely Large Telescopes and
future mission concepts
like Origins,
HabEx, and LUOVIR.
ACKNOWLEDGEMENTS
The authors acknowledge funding from the
Brinson Foundation. and the Carl Sagan
Institute
REFERENCES
Arney G, Domagal-Goldman SD, Meadows VS, et
al. 2016. Astrobiology 16(11):873 -- 99
Barstow JK, Irwin PGJ. 2016. MNRS Lett.
461(1):L92 -- 96
Batalha, N. M., Rowe, J. F., Bryson, S. T., et al.
2013, ApJS, 204, 24
Battersby, C., Armus, L., Bergin, E., et al. 2018,
Nature Astronomy, 2, 596
train
Berger, T. A.; Huber, D., Gaidos, E. & van
Saders, J. L., ApJ, Volume 866, 2, 2019
Betremieux Y, Kaltenegger L. 2013. Ap. J. Lett.
772:L31
Betremieux Y, Kaltenegger L. 2014. Ap. J. 791:7
Borucki, W. J., Agol, E., Fressin, F., et al. 2013,
Sci, 340, 587
Borucki, W. J., Koch, D. G., Basri, G., et al. 2011,
ApJ, 736, 19
Claire, M. W., Sheets, J., Cohen, M., et al. 2012,
ApJ, 757, 95
Ehrenreich D, Tinetti G, Lecavelier des Etangs
A, et al. 2006. Astron. Astrophys. 448:379
Garcıa Munoz A, Zapatero Osorio MR, Barrena
R, et al. 2012. Ap. J. 755(2):103
Haqq-Misra JD, Saxena P, Wolf ET, Kopparapu
RK. 2016. Ap. J. 827:2
Hedelt P, von Paris P, Godolt M, et al. 2013.
Astron. Astrophys. 553:A9
Holland, R. 2006, Philos. Trans. R. Soc. London
B, 361, 903
Johns, D., Marti, C., Huff, M., McCann, J,
Wittenmyer, R.A., Horner, J., and Wright,
D.J., The Astrophysical Journal Supplement
Series, 239, 1 (2018)
Kaltenegger L, Henning WG, Sasselov DD. 2010.
Astron. J. 140:1370 -- 80
Kaltenegger L, Madden J., Lin Z., Rugheimer S.,
Segura A., Luque R., Palle E., Espinoza N.,
2019, The Habitability of GJ 357d: Possible
Climate and Observability, ApJL, 883, 2
Kaltenegger L, Sasselov DD. 2010. Ap. J.
708:1162 -- 67
Kaltenegger L, Sasselov DD. 2011. Ap. J. Lett.
736(2):L25
Kaltenegger L, Traub WA, Jucks KW. 2007.
Astron. J. 658(1):598 -- 616
Kaltenegger L, Traub WA. 2009. Ap. J. 698:519
Kaltenegger, L. 2017, Annual Review of
Astronomy and Astrophysics, 55, 433
CE. 2014. PNAS 111(35):12641 -- 46
1383
Kasting JF, Kopparapu R, Ramirez RM, Harman
Kasting, J. F., & Ackerman, T. P. 1986, Sci, 234,
Lederberg J. 1965. Nature 207(4992):9 -- 13
Lin. Z. & Kaltenegger, L., High-resolution
reflection spectral models of Proxima-b and
Trappist-1e, MNRAS, published online DOI:
10.1093/mnras/stz3213
Lippincott ER, Eck RV, Dayhoff MO, Sagan C.
1967. Ap. J. 147:753
Lovelock JE. 1965. Nature 207(997):568 -- 70
Lyons, T.W., Reinhard, T.W., Planavsky, N.J.,
Nature volume 506, pages307 -- 315(2014)
OMalley-James & Kaltenegger L, 2019, Lessons
from Early Earth: UV surface radiation
should not limit the habitability of active M
star systems, MNRAS, 485 (4), 5598-5603
Marty, B., Zimmermann, L., Pujol, M., Burgess,
R., & Philippot, P. 2013, Sci, 342, 101
Mennesson, B., Gaudi, S., Seager, S., et al. 2016,
of
in
Photo-Optical
Society
Engineers
Instrumentation
(SPIE)
Conference Series, Vol. 9904, Proc. SPIE,
99040L, Ser., ed. M Mendillo, A Nagy,
JHWaite, pp. 369 -- 80. Washington, DC: Am.
Geophys. Union
Misra A, Meadows V, Claire M, Crisp D. 2014.
Astrobiology 14:67
Niro, F., Jucks, K., & Hartmann, J.-M. 2005a,
JQSRT, 95, 469
Niro, F., Von Clarmann, T.,
Jucks, K., &
Hartmann, J.-M. 2005b, JQSRT, 90, 61
O'Malley-James JT, Kaltenegger L. 2017b. MNRS
Lett. 469:L26 -- 30
Palle E, Osorio MRZ, Barrena R, et al. 2009.
Nature 459:814 -- 16
Pavlov AA, Kasting JF, Brown LL, et al. 2000. J.
Geophys. Res. 105:11981 -- 90
Quintana, E. V., Barclay, T., Raymond, S. N., et al.
2014, Sci, 344, 277
Rauer H, Gebauer S, Paris PV, et al. 2011.
Astron. Astrophys. 529(5):A8
Rodler F, Lopez-Morales M. 2014. Ap. J. 781:54
Rugheimer S, Kaltenegger L, Zsom A, et al.
2013. Astrobiology 13:251 -- 69
Rugheimer, S., & Kaltenegger, L. 2018, ApJ, 854,
19
Sagan C, Thompson, W.R., Carlson R, et al.
1993. Nature 365(6448):715 -- 21
Schwieterman, E. W., Cockell, C. S., & Meadows,
V. S. 2015, Astrobiology, 15, 341
Segura A, Kasting JF, Meadows V, et al. 2005.
Astrobiology 5:706 -- 25
Segura A, Meadows VS, Kasting JF, et al. 2007.
Astron. Astrophys. 472(2):665 -- 79
Sharpe, S. W., Johnson, T. J., Sams, R. L., et al.
2004, ApSpe, 58, 1452
Snellen I, de Kok R, Le Poole R, et al. 2013. Ap. J.
764:182
Som, S. M., Catling, D. C., Harnmeijer, J. P.,
Polivka, P. M., & Buick, R. 2012, Natur, 484,
359
Stevenson, K. B., Lewis, N. K., Bean, J. L., et al.
2016, Publications of the Astronomical
Society of the Pacific, 128, 094401
The LUVOIR Team. 2018, arXiv e-prints,
arXiv:1809.09668
Torres, G., Kipping, D. M., Fressin, F., et al. 2015,
ApJ, 800, 99
Traub, W. A., & Stier, M. T. 1976, ApOpt, 15, 364
Traub, W.A., Jucks K. 2002. In Atmospheres in
Aeronomy,Vol. 130, Geophys. Monogr.
Udry, S., Bonfils, X., Delfosse, X., et al. 2007,
A&A, 469, L43
Comparative
System:
the
Solar
Vidal-Madjar A, Arnold A, Ehrenreich D, et al.
2010. Astron. Astrophys. 523:A57
Wagner, G., & Birk, M. 2003, JQSRT, 82, 443
Walker JCG. 1977. Evolution of the Atmosphere.
New York: Macmillan
Zahnle K, Arndt N, Cockell C, et al. 2007. Space
Sci. Rev. 129(1 -- 3):35 -- 78
|
1302.4621 | 1 | 1302 | 2013-02-19T14:38:15 | Secular Light Curves of Comets C/2011 L4 Panstarrs and C/2012 S1 ISON Compared to 1P/Halley | [
"astro-ph.EP"
] | We used 1329 observations of comet C/2011 L4 Panstarrs and 1897 observations of C/2012 S1 ISON to create the Secular Light Curves (SLCs), and to compare them to comet 1P/Halley (3172 observations). These comets are at the present time approaching perihelion. Thus the information contained in this report is preliminary. | astro-ph.EP | astro-ph | 1
C:/comets/C2011L4/SLC L4+S1/130218
“Secular Light Curves of comets
C/2011 L4 Panstarrs and
C/2012 S1 ISON
Compared to 1P/Halley”
Ignacio Ferrín,
Institute of Physics,
Faculty of Exact and Natural Sciences,
University of Antioquia,
Medellín, Colombia, 05001000
[email protected]
Submitted for publication
2
Abstract. We used 1329 observations of comet C/2011 L4 Panstarrs and
1897 observations of C/2012 S1 ISON to create the Secular Light Curves
(SLCs), and to compare them to comet 1P/Halley (3172 observations), the
famous comet of 1986. These comets are at the present time approaching
perihelion. Thus the information contained in this report is preliminary but
serves for planning purposes. The scientific results found in this investigation
are:
(1) Comets C/2011 L4 and C/2012 S1 turned on farther away than comet
1P/Halley at R= - 6.2±0.1 AU. Since water ice can not subl imate at distances
R< -6 AU, these comets must be rich in substances more volatile than water,
like CO or CO2.
(2) C/2011 L4 exhibits a “Slowdown Distance”, SD, at which the brightness
increase rate slows down to a more relaxed pace. The brightness law brakes.
This is reminiscent of the same process for comet 1P/Halley and 11 other
comets listed in Table 1. We find R(SD) = -3.6±0.1 AU, m(SD)= 7.1±0.1, while
for comet 1P/Halley R(SD) = -1.7±0.1 AU, m(SD)= 5.6±0.1.
(3) The absolute magnitude of comet C/2011 L4 is m(Δ,-R) = m(1,-1)=+5.2
compared with m(1,-1)= +3.9 for comet 1P/Halley. After passing the SD, the
comet is increasing its brightness with a shallow power law R+1.25. We predict a
magnitude at perihelion m(q)= -2.4±0.4, but the comet will be in the Sun’s glare.
(4) Comet C/2012 S1 ISON poses a dilema about its future luminosity behavior :
(A) On the one hand Figure 5 shows evidence of a future event of SD. In this
case the future brightness is uncertain and we have to wait until the SD event
happens. We assign a probability of 25% to this possibility. (B) On the other
hand, Figure 4 shows evidence that this comet had a recent “Slowdown
Distance Event”, maybe at R(SD) = -7.2±0.1 AU, m(SD)= +12.2±0.1. In this
case the power law keeps its pace, and then the comet will be at magnitude
m(1,q) = -16±2 at perihelion, brighter than the full moon. Because the orbit
enters the Roche Limit, and the calculated temperature at perihelion is 2919º K,
the comet may desintegrate giving an awesome cosmic spectacle. We assign
to this possibility a probability of 75%.
(5) The photometric age, P-AGE, of the comets has been estimated. We find
that for both comets P-AGE < 3 comet years which corresponds to “baby
comets”.
(6) The m(SD) vs R(SD) plot shows the Jupiter family comets concentrated in a
small area of this phase space. The reason for this is not understood. Some
plots shown in this work exhibit complexity beyond current understanding.
3
1. Introduction
2013 offers the opportunity to observe two new comets coming from the
Oort cloud, C/2011 L4 Panstarrs and C/2012 S1 ISON. Both have orbits with
excentricity e = 1.0.
In June 6th, 2011, astronomers at the Institute of Astronomy of the
University of Hawaii, discovered a comet with the designation C/2011 L4
Panstarrs (Wainscoat and Micheli, 2011).
We will show that when the comet reached to R= -3.6 AU (the minus sign
indicates pre-perihelion location), the brightness law slowed down, from R +8.9 to
R+1.25 . We call this the Slowdown Distance, SD. This is not surprising. 12
other comets, including comet 1P/Halley, have exhibited such a SD (Ferrín,
2010) and they are listed in Table 1. The coordinates of the SD are R(SD)= -
3.6±0.1 AU, magnitud m(SD)= +7.1±0.1 (see Figure 1).
In September 21st of 2012, Nensky and Novichonok discovered comet
C/2012 S1 ISON at the notable distance of -6.3 AU (Nevsky and Novichonok,
2012). The object is coming from the Oort cloud with an excentricity of
1.0000013, esentially parabolic. It has been increasing in brightness at a rate
R+4.35 and if it were to continue at this rate, it would certainly attain a
magnitude much brighter than the full moon. However we will show that this
possibility has only a 75% chance of being correct. There is a 25% pro bability
that the comet will exhibit a SD event and attain a much smaller brightness (see
Figures 2-5).
In the phase space m(SD) vs R(SD) (see Figure 5) 10 comets lie in a
rather small area of the phase space, and two comets C/1995 O1 Hale-Bopp,
and C/2011 L4 Panstarrs, lie at the center and left of the diagram. There is no
physical explanation of why so many comets lie in such a narrow area of this
phase space or why there should a SD. However , a hypothesis is advanced in
the text.
2. Data sets
The data sets to create the secular light curves are available in internet:
(1) The Cometary Science Archive http://www.csc.eps.harvard.edu/index.html
is a site to visit because it contains useful scientific information of current
and past comets.
(2) Another usefull site is the Minor Planet Center repository of astrometric
observations, http://www.minorplanetcenter.net/db_search .
(3) Seiichi Yoshida’s web place http://www.aerith.net contains many raw light
curves.
(4) The Yahoo site http://tech.groups.yahoo.com/group/CometObs/ contains up
to date observations by many observers for many comets.
(5) The spanish group measures magnitudes with several CCD apertures:
http://www.astrosurf.com/cometas-obs. This allows the determination of
4
Infinite aperture magnitudes (Ferrín, 2005b), which better represent the
whole magnitude of a comet.
(6) The site http://www.observatorij.org/cobs/ contains many observations in
the ICQ format, as well as news concerning comets, and it is mantained
by the Crni Vrh Observatory.
(7) The german comet group publishes their observations at
http://kometen.fg-vds.de/fgk_hpe.htm .
In all we used 1329 observations of comet C/2011 L4 Panstarrs and
1897 observations of C/2012 S1 ISON to create the Secular Light Curves
(SLCs), and compared with 3172 observations of comet 1P/Halley.
3. SLCs
The elaboration of the Secular Light Curves (SLCs) follows closely the
procedures described in the “Atlas of Secular Light Curves of Comets” (Ferrín,
2010). The preferred phase space to describe the luminous behaviour is the
m(1,R) = m( Δ ,R) – 5 log Δ vs Log R plane, where Δ is the Sun -Earth distance
and R the heliocentric distance. In the m(1,R) vs Log R diagram, powers of R,
R+n, plot as straight lines of slope 2.5 n. The R+n behavior is easy to spot and
measure.
In this work we adopted the “envelope of the data set” as the correct
interpretation of the observed brightness. There are many physical effects that
affect comet observations like twilight, moon light, haze, cirrus clouds, dirty
optics, lack of dark adaptation, and in the case of CCDs, sky background too
bright, insufficient time exposure, insufficient CCD aperture error. All these
effects diminish the captured photons emitted by the comet and the observer
makes an error downward, toward fainter magnitudes. There are no
corresponding physical effects that could increase the perceived brightness of a
comet. Thus the “envelope” is the correct interpretation of the data. In fact the
envelope is rather sharp, while the anti -envelope is diffuse and irregular.
C/2011 L4 Panstarrs (Figure 1). To help in the interpretation of the comet’s
lightcurve, we will compare with the observed SLC of comet 1P/Halley (Ferrín,
2010). The Figure shows that C/2011 L4 Panstarrs turned on much before
comet Halley, at -R < -10 AU. Since water can not sublimate at distances
beyond -6 AU, the comet must be composed of something more volatile than
water, probably CO or CO2. Spectroscopy could tell.
Comet 1P/Halley exhibited a well behaved SLC with a notable
“Slowdown Distance” after which the rate of brightness increase, dimin ished.
There was a break in the linear law. For 1P the slowdown distance was
R= -1.7±0.1 AU. And the deceleration took place at a magnitude of m=
+5.6±0.1.
For C/2011 L4 Panstarrs before the SD, the power law was R+8.9 and
after it is R+1.25. Figure 1 gives the absolute magnitude (at Δ=R=1AU) m(1, -
1)=+5.2. The magnitude of the comet can then be predicted
5
m(Δ, R) = m(1,-1) + 5 log Δ + 2.5 n Log R
(1)
This formula predicts a magnitude at perihelion, m(q)= m(1.109, 0.30) =
-2.4±0.4 . The comet should be as bright at the planet Venus but difficult to
see due the the Sun’s glare.
C/2012 S1 ISON (Figures 2-5). Comparison will also be made to comet
1P/Halley. The comet exhibits a shallow brightness increase with a power law
is R+4.35 (Figure 2). If this rate were to continue up to perihelion, the comet
would indeed be much brighter than the full Moon. This comet poses a dilema:
(A) On one hand, Figure 5 shows evidence of a future event of SD: The comet
is approaching the Oort Cloud region of comets. In this case after the SD
event, the new power law will allow a definitive prediction of the brigtness of the
comet. The predicted location of the SD is shown in Figure 4. The SD event
has to take place before October 21st, moment at which the comet reaches to
1.25 AU from the Sun, the smallest distance at which a SD event has taken
place. For example in that case the comet could pass from a R +4.35 power law
to the R+1.25 very relaxed pace of comet C/2011 L4. This is not very probable,
however, because such a shallow power law is very uncommon. Additionally,
the two comets are not dynamically related. Thus we assign a probability of
25% to this case.
(B) On the other hand, Figure 4 shows strong evidence that this comet had a
recent “Slowdown Distance Event”. If we assume the same slope of the SLC as
that of comet Halley, then the event could have taken place at R(SD) = -7.2±0.1
AU, m(SD)= +12.2±0.1 (see Figure 2, bottom left corner). Then the current
power law is the final law after the SD and it should keeps its pace. A few lines
below we calculate that the comet will be at magnitude m(1,q) = -16±2 at
perihelion, brighter than the full moon. We assign a probability of 75% to this
case.
Using T = 325º/SQRT(d), d = q = 0.0124 AU, we find T (q) = 2919º K. It
is questionable if the comet is going to survive such a temperature.
It is possible to calculate the Roche Limit for the Sun, using a density of 1
gm/cm3 for the comet. We find R(Roche)= 1.92 10 6 km. While the comet
reaches perihelion at q= 0.012472 AU = 1.87 10 6 km. Thus the comet enters
the Roche Limit of the Sun. The high temperature and the Roche Limit suggest
that this comet is going to desintegrate in a cosmic spectacle.
Figure 3 gives the absolute magnitude (at Δ=R=1AU) m(1, -1)=+4.6.
Using R+4.35 the magnitude of the comet can then be predicted
m(Δ, R) = m(1,-1) + 5 log Δ + 2.5 n Log R
(2)
= 4.6 + 5 log Δ + 10.9 Log R
This formula predicts a magnitude at perihelion m(q) = m(0.99, 0.01) =
6
P-AGE [comet years] = 1440 / [ Asec Rsum ]
-16±2, brighter than the full moon.
4. Photometric Age, P-AGE
The photometric age of a comet is a proxy for age and may be defined as
(Ferrín, 2010):
where Asec = amplitude of the secular light curve = V(1,1,0) – m1(1,1), and
Rsum is the sum of the turn on, Ron, and turn off, Roff, distances o f the comet.
P-AGE is measured in comet years to be sure that they are not confused with
current years. The constant is chosen so that comet 29P/Neu jmin 1 has a
P-AGE = 100 cy. For both comets in this work we do not know any of these
parameters so it would seem impossible to determine P-AGE. However in
Figure 6 we show a correlation P-AGE vs Ron. Since we know that Ron < -10
AU for both comets, Figure 6 returns P-AGE < 3 cy. In the classification of
Ferrín (2010) these are “baby comets” ( those with P-AGE < 4 cy ).
5. m(SD) vs R(SD) diagram.
(3)
This diagram is shown in Figure 5 and it is based on the data presented
in Table 1. It shows 10 comets closely located in a small a rea of this phase
space, and two comets beyond their area. Most of these comets are members
of the Jupiter family but notice that 4 members of the Oort Cloud are located
inside the Jupiter Family Box. The reason why Oort Cloud and JF comets are
segregated in this diagram, is not understood.
6. Why should there be a SD?
When a comet from the Oort cloud (e=1.0) falls into the Sun, the upper
layer of the nucleus contains fresh volatiles like CO, CO2 an H2O. As the
comet approaches, temperature increases, and the first one to sublimate is CO.
Next sublimates CO2 and finally H2O. This is due to their vapor pressure.
Sublimating CO or CO2, the light curve far from the Sun is a straight line with a
power law ~ R+9.1±2.0 (Table 2). The H2O does not have sufficient temperature
to sublimate and thus CO or CO2 control the surface sublimation and the light
curve. The sublimation rate increases as the comet approaches the Sun, and
at a given temperature, H2O overpowers CO or CO2, and H2O now controls
the surface. The brightness increase decreases its rate according to the new
sublimation rules. This is a hypothetical mechanism behind the SD.
However, there have been no numerical or theoretical models to explain this
discontinuity, and thus the hypothesis remains unconfirmed.
temperature.
to a
R(SD) is a distance and it must be related
Observationally (Figure 5) we find a range 1.24 < R(SD) < 2.09. Using T =
325º/SQRT(d), we find 222º < T < 289º K. This means that the comets in our
data base have had their events within a range of temperatures of only 67ºK,
perhaps due to different pole orientations.
7
m(SD) would seem to be related to the ratio CO/H2O or CO2/H2O. It is
expected that espectroscopic and compositional estudies of C/2011 L4
Panstarrs and C/2012 S1, would reveal differences that can be correlated to the
list of SD’s properties compiled in Table 1 including JF Comets.
7. Conclusions
(1) We present the secular light curves, SLCs, of comets C/2011 L4 Panstarrs
and C/2012 S1 ISON. Both comets turned on beyond -10 AU from the Sun.
For comparison comet 1P/Halley turned on at R= - 6.2±0.1 AU. Since water
ice can not sublimate at distances R<-6 AU, these comets have to contain
substances more volatile than water, like CO or CO2.
(2) We measure the Slowdown Distance of C/2011 L4 Panstarrs. This is the
distance at which the brightness increase rate slows down to a more relaxed
pace. The brightness law brakes. This is reminiscent of the same process for
comet 1P/Halley and 11 other comets listed in Table 1. We find R(SD) =
-3.6±0.1 AU, m(SD)= 7.1±0.1, while for comet 1P/Halley R(SD) = -1.7±0.1 AU,
m(SD)= 5.6±0.1.
(3) We derive the absolute magnitude of C/2011 L4 Panstarrs and the power
laws that define its brightness behavior. The absolute magnitude is m(Δ, -R) =
m(1,-1)=+5.2 compared with m(1,-1)= +3.9 for comet 1P/Halley. After passing
the SD, the comet is increasing its brightness with a shallow power law R +1.25.
With this information the magnitude at perihelion can be calculated and we find
m(q)= -2.4±0.4 .
(4) Comet C/2012 S1 ISON poses a dilema about its future luminosity behavior:
(A) On the one hand Figure 5 shows evidence of a future event of SD. In this
case the future brightness is uncertain and we have to wait until the SD event
happens. We assign a probability o f 25% to this possibility. (B) On the other
hand, Figure 4 shows evidence that this comet had a recent “Slowdown
Distance Event”, maybe at R(SD) = -7.2±0.1 AU, m(SD)= +12.2±0.1. In this
case the power law keeps its pace, and then the comet will be at magnitude
m(1,q) = -16±2 at perihelion, brighter than the full moon. Because the orbit
enters the Roche Limit, and the calculated temperature at perihelion is 2919º K,
the comet may desintegrate giving an awesome cosmic spectacle. We assign
to this possibility a probability of 75%.
(5) The photometric age, P-AGE, of the comets has been estimated. We find
that for both comets P-AGE < 3 comet years which corresponds to “baby
comets”.
(6) The m(SD) vs R(SD) plot shows the Jupiter family comets concentrated in a
small area of this phase space. The reason for this is not understood. Some of
the plots shown in this work, exhibit complexity beyond current understanding.
8
References
Ferrín, I., 2005a. Secular Light Curve of Comet 28P/Neujmin 1, and of Comets
Targets of Spacecraft, 1P/Halley, 9P/Tempel 1, 19P/Borrelly, 21P/Grigg-
Skejellerup, 26P/Giacobinni -Zinner, 67P/Chruyumov-Gersimenko,
81P/W ild 2. Icarus 178, 493-516.
Ferrín, I., 2005b. Variable Aperture Correction Method in Cometary
Photometry, ICQ 27, 249-255.
Ferrín, I., 2010. Atlas of Secular Light Curves of Comets. PSS, 58, 365 -391.
Available in color and lanscape format at
http://arxiv.org/ftp/arxiv/papers/0909/0909.3498.pdf
Green, D.W ., 2013. Cometary Science Archive.
http://www.csc.eps.harvard.edu/index.html .
Nevsky, V., Novichonok, A., 2012. Comet C/2012 S1 ISON. CBET 3228.
Wainscoat, R., Micheli, M., 2011. Comet C/2011 L4 (PANSTARRS).
IAUC 9215.
9
Figure 1. The secular light curve of comet C/2011 L4 compared with that of
comet 1P/Halley. A number of features can be discerned. The turn on
distance of C/2011 L4 is much farther away than that of 1P, in fact beyond 10
AU. Since water cannot sublimate a R>6 AU the comet has to be made of
something more volatile than water, like CO or CO2. Both comets exhib it a
“slowdown distance”. For 1P it is located at R= -1.7±0.1 AU, while for C/2011
L4 it is located at R= -3.6±0.1 AU. For comparison comet C/1995 O1 Hale-
Bopp had R= -6.4±0.1 AU. The absolute magnitudes of 1P and C/2011 L4 are
m(1,-1)= +3.9 and +5.2, but C/2011 L4 exhib its a very low magnitude increase
R+1.25 vs comet 1P/Halley R+3.35 . Curiously both comets exhibit the same
slope before SD. The reason why we select the envelope as the correct
interpretation of the light curve, is explained in the text. We used 1329
observations of the comet to create the SLC. For comet Halley we used 3172
observations.
10
Figure 2. Enlarged region around the current time location of comet C/2012 S1
with up to date observations (2013 Feb 16 th). The brightness is increasing with
a power law R+4.35 (Table 1). This shallow value makes it probable that the
comet has already passed its SD phase. In this case the current rate serves as
a predictor. Or the comet will experience a SD event, in which case the rate
after the event will serve as a predictor. For comet Halley we used 3172
observations. For comet C/2012 S1 1897 observations.
11
Figure 3. The secular light curve of comet C/2012 S1 compared with that of
comet 1P/Halley. A number of features can be discerned. T he turn on
distance of C/2012 S1 is much farther away than that of 1P. Since water
cannot sublimate a R> -6 AU, the comet has to be composed of something
more volatile than water, like CO or CO2. We find the absolute magnitude
m(1,1)= +4.6±0.2. Comet Halley exhib its a “slowdown distance” located at
R(SD)= -1.7±0.1 AU, m(SD)= +5.6±0.1. For comet C/2012 S1 we have two
possib ilities:
(A) The comet has not yet arrived to its SD event, as suggested by Figure 5. In
which case the final brightness is too early to tell and we have to wait until past
the SD event.
(B) If we assume a slope identical to those of comet 1P/Halley and C/2011 L4,
that is R+8.92 , then comet C/2012 S1 may have had a SD event at R = -7.2±0.1
AU, m(1,R)= +12.2±0.1 (lower left part of the diagram) , in which case the
observed power law is the power law after SD event. Figure 4 also suggests
very strongly that the shallow power law exhib ited by th is comet is in accord
with that exh ib ited by Oort Cloud comets, after SD events (confirmation Table
2). In this case the comet would reach to magnitude m(1,q)=-17 at perihelion,
brighter than the full Moon , if the pace is sustained, as has been sustained in
previous cases.
12
Figure 4. A m(1,1) vs n Diagram for Comets. Or the Pre-SD vs Post-SD vs
Oort Cloud vs JF comets diagram. The values measured in Table 2 are plotted
here. They separate Pre and Post-SD events, and Oort Cloud Comets vs
Jupiter Family comets into 4 distinct sectors. In particular the Pre-Post SD of
Oort Cloud comets are clearly separated. The location of comet C/2012 S1 is
nearer to Oort Cloud Comets in a Post-SD Phase. In this case the current
power law, R+4.35 , is a good estimate of its future brightness.
13
Figura 5. The Slowdown Magnitudes of 13 comets are plotted vs their
Slowdown Distances from Table 1. 10 of 12 (83%) lie in a narrow vertical zone
centered at R = -1.7 AU. Another comet (C/2011 L4) lies in the m iddle of the
diagram, and another (C/1995 O1 Hale-Bopp) lies to the left. The location of
the Slowdown Distance of comet C/2012 S1 in ind icated in two regions: (A) if
the event already passed, down and to the left. (B) If the event is going to
happen, to the right. Notice that the comet is moving toward the region of other
Oort Cloud comets. This favors option (B) (that the SD event has not
happened). However Figure 4 favors the opposite, that the event has taken
place. Table 1 allows us to define the Jupiter Family Box of Comets, as 1.24 <
R(SD) < 2.09 AU, and 4.82 < m(SD) < 13.81. Inside this box there are 4 Oort
Cloud comets. Up to this moment there is no physical interpretation of why so
many comets should lie in such a small area of this phase space, although a
preliminary hyphothesis is advanced in the text.
14
Figure 6. The correlation between the turn on point Ron vs the pho tometric
age P-AGE measured in comet years. The correlation shows that comets with
turn on points beyond 10 AU are less than 3 comet years old. Older comets
have to get nearer to the Sun to get activated. The two comets studied in th is
work, turned on beyond 10 AU and thus are classified as “baby comets” (Ferrín,
2010).
15
Table 1. Distances, magnitudes at Slowdown,q,e,i,Tiss.
----------------------------------------------------------------
R(SD) m(SD) q e i Tiss Comet
----------------------------------------------------------------
-1.8 8.9 0.316 1.000240 119.9 0.00 C/1956 R1 Arend-Roland
-3.6 7.1 0.302 1.000028 84.3 0.00 C/2011 L4 Panstarrs
-1.56 7.55 0.171 1.000018 77.1 0.00 C/2006 P1 McNaught
---- ---- 0.012 1.000004 62.1 0.00 C/2012 S1 ISON
-1.39 7.6 0.099 0.999903 81.7 0.06 C/2002 V1 NEAT
-1.7 6.73 0.230 0.999758 124.9 -0.33 C/1996 B2 Hyakutake
-6.29 4.82 0.914 0.994929 89.4 0.04 C/1995 O1 Hale-Bopp
-1.68 5.55 0.586 0.967813 162.3 -0.61 1P/Halley
----------------------------------------------------------------
-1.57 10.6 1.031 0.707045 31.9 2.46 21P/Giacobinni-Zinner
-1.24 9.38 1.059 0.694533 13.6 2.64 103P/Hartley 2
-1.81 16.24 1.057 0.659295 11.7 2.81 46P/Wirtanen
-1.9 10.0 1.598 0.537385 3.2 2.88 81P/Wild 2
-2.09 13.81 1.509 0.516946 20.5 2.90 9P/Tempel 1
----------------------------------------------------------------
Table 2. Absolute magnitudes and power laws.
Comet
m(1,1) n in R^n
-------------------------------------------------------
Oort Cloud Comets Pre SD Post SD
-------------------------------------------------------
+7.2 and +4.0
+6.3±0.1
C/1956 R1 AR
+8.9 and +1.252
+5.2±0.1
C/2011 L4 Pan
C/2006 P1 McN
+5.2±0.1 +10.6 and +1.55
C/2012 S1 ------ ---- --- +4.32
+6.7±0.1 +13.1 and +2.18
C/2002 V1 NEAT
+4.8±0.1 +11.6 and +2.33
C/1996 B2 Hy
C/1995 O1 HB
-0.6±0.1 +10.7 and +2.58
1P/Halley
+8.9 and +3.35
+3.9±0.1
-------------------------------------------------------
Jupiter Family
-------------------------------------------------------
+9.1 and +5.16
+8.0±0.1
21P
+9.5 and +5.55
+8.3±0.1
103P
46P
+7.6±0.1
+5.2 and +7.20
+9.3 and +7.03
+5.8±0.2
81P
9P
+6.4±0.2
+7.7 and +6.50
-------------------------------------------------------
1 Comet C/2011 L4 has the smallest Post SD n value
of the whole sample.
2 C/2012 S1 has a small rate of brightness increase
reminiscent of Oort Cloud Comets Post SD. This
data is plotted in Figure 4.
|
1911.00816 | 1 | 1911 | 2019-11-03T03:17:35 | A Dynamic Trajectory Fit to Multi-Sensor Fireball Observations | [
"astro-ph.EP"
] | Meteorites with known orbital origins are key to our understanding of Solar System formation and the source of life on Earth. However, these pristine samples of space material are incredibly rare. Less than 40 of the 60,000 meteorites held in collections around the world have known dynamical origins. Fireball networks have been developed globally in a unified effort to increase this number by using multiple observatories to record, triangulate, and dynamically analyse ablating meteoroids as they enter our atmosphere. The accuracy of the chosen meteoroid triangulation method directly influences the accuracy of the determined orbit and the likelihood of possible meteorite recovery.
There are three leading techniques for meteoroid triangulation discussed in the literature: the Method of Planes, the Straight Line Least Squares method, and the Multi-Parameter Fit method. Here we describe an alternative method to meteoroid triangulation, called the Dynamic Trajectory Fit. This approach uses the meteoroid's 3D dynamic equations of motion to fit a realistic trajectory directly to multi-sensor line-of-sight observations. This method has the ability to resolve fragmentation events, fit systematic observatory timing offsets, and determine mass estimates of the meteoroid along its observable trajectory.
Through a comprehensive Monte-Carlo analysis of over 100,000 trajectory simulations, we find this new method to more accurately estimate meteoroid trajectories of slow entry events ($<$25\,km/s) and events observed from low convergence angles ($<$10$^{\circ}$) compared to existing meteoroid triangulation techniques. Additionally, we triangulate an observed fireball event with visible fragmentation using the various triangulation methods to show that the proposed Dynamic Trajectory Fit implementing fragmentation to best match the captured multi-sensor line-of-sight data. | astro-ph.EP | astro-ph |
A Dynamic Trajectory Fit to Multi-Sensor Fireball
Observations
Trent Jansen-Sturgeon∗, Eleanor K. Sansom∗, Hadrien A. R. Devillepoix∗,
Philip A. Bland∗, Martin C. Towner∗, Robert M. Howie∗, Benjamin A. D.
Hartig∗
Abstract
Meteorites with known orbital origins are key to our understanding of Solar
System formation and the source of life on Earth. However, these pristine
samples of space material are incredibly rare. Less than 40 of the 60,000 me-
teorites held in collections around the world have known dynamical origins.
Fireball networks have been developed globally in a unified effort to increase
this number by using multiple observatories to record, triangulate, and dy-
namically analyse ablating meteoroids as they enter our atmosphere. The
accuracy of the chosen meteoroid triangulation method directly influences
the accuracy of the determined orbit and the likelihood of possible meteorite
recovery.
There are three leading techniques for meteoroid triangulation discussed
in the literature: the Method of Planes, the Straight Line Least Squares
method, and the Multi-Parameter Fit method. Here we describe an alterna-
tive method to meteoroid triangulation, called the Dynamic Trajectory Fit.
This approach uses the meteoroid's 3D dynamic equations of motion to fit
a realistic trajectory directly to multi-sensor line-of-sight observations. This
method has the ability to resolve fragmentation events, fit systematic obser-
vatory timing offsets, and determine mass estimates of the meteoroid along
its observable trajectory.
Through a comprehensive Monte-Carlo analysis of over 100,000 trajec-
tory simulations, we find this new method to more accurately estimate mete-
oroid trajectories of slow entry events (<25 km/s) and events observed from
low convergence angles (<10◦) compared to existing meteoroid triangulation
techniques. Additionally, we triangulate an observed fireball event with visi-
ble fragmentation using the various triangulation methods to show that the
∗School of Earth and Planetary Sciences, Curtin University, GPO Box U1987, Perth,
WA 6845, Australia
Preprint submitted to Astronomical Journal
November 5, 2019
proposed Dynamic Trajectory Fit implementing fragmentation to best match
the captured multi-sensor line-of-sight data.
Keywords: Fireball, Network, Triangulation, Meteoroid, Dynamics
1. Introduction
Fireball networks have been around since the 1960's with the specific
goal of observing meteors from multiple stations to determine their past and
future trajectories (Ceplecha, 1961). The meteoroids of real interest are the
bright, deeply penetrating kind, with the highest chance of surviving the
violent atmospheric entry process to produce meteorites. Finding meteorites
with known orbits is key for giving these cosmic samples a regional context in
the greater Solar System, potentially helping to answer some of the biggest
questions in planetary science, such as the Solar System formation, and the
origin of life on Earth. As of mid-2019, only 36 out of about 60,000 collected
meteorites have known orbits; 5 of which have been found in recent history
by the Global Fireball Observatory (GFO), a global collaboration of fireball
networks, including Australia's Desert Fireball Network (DFN).
Successful recovery of the incoming meteorite requires accurate knowledge
of the fall position. If found, it is highly desirable to have a well constrained
and accurate orbit associated with the sample. Determining an orbit re-
quires the entry radiant and velocity of the meteoroid, while prediction of
fall positions require darkflight modelling, where darkflight is the period of
meteoroid free fall to Earth after visible observations cease, during which
the body is strongly influenced by its size and shape, as well as atmospheric
winds. At the heart of all this dynamic analysis lies the triangulation and
modelling of the observed luminous trajectory; giving both the darkflight
and orbit determination their initial conditions. To improve the accuracy of
these predictions, we must first improve the accuracy of our triangulation
modelling techniques.
Three prominent methods of meteoroid triangulation have been docu-
mented and used in the past; Method of Planes (Ceplecha, 1987), Straight
Line Least Squares (Borovicka, 1990), and Multi-Parameter Fit (Gural,
2012). These three methods are outlined conceptually below. For more
detail and mathematical rigour, please refer to their respective papers.
A notable additional technique is the particle filter modelling method of
Sansom et al. (2019) as an alternative to the traditional triangulation meth-
ods. While particle-type approaches are thorough, they are also quite com-
putationally intense and are not feasible as the default triangulation method
2
for large meteoroid data-sets. Instead, it is generally best suited to special
cases when a surviving meteorite is suspected.
1.1. Method of Planes (Ceplecha, 1987)
Although the Method of Planes (MOP) is the oldest and least accurate of
the three prominent triangulation methods, it is very computationally simple
and often used for constructing the initial trajectory guess for more complex
methods, such as the Straight Line Least Squares and Multi-Parameter Fit
(see Section 1.2 and Section 1.3). MOP comprises four main steps: plane
construction, radiant formation, position determination, and velocity fitting.
To begin, MOP constructs a plane for every sensor. This plane includes
the sensor's location and best fits to its associated observation rays using a
least squares approach. It does so by adjusting the plane normal to minimise
the square of the angular residuals between the rays and the plane.
Once the optimum plane is calculated for each sensor, they are intersected
in 3D space to determine the straight line trajectory.
In the case where
more than two sensors recorded the meteoroid, a statistical weighting can be
used to combine the straight line solutions from every different sensor-pair
combination to produce one unique straight-line trajectory. This weighting
is based on the convergence angle between the two planes as well as on the
combined angular span of the observed meteoroid across the sensors.
Positions along the determined straight-line trajectory are found for every
observation (regardless of time) as the closest point on the trajectory-line
from that observed line-of-sight. These 3D positions are generally calculated
by the intersection of the trajectory itself with a series of planes that each
contain an individual line-of-sight and its associated optimum plane normal.
Lastly, the velocities are determined by fitting a model to the positional
lengths along the trajectory as a function of time. These velocity models and
fitting methods are described by Pecina and Ceplecha (1983, 1984). However,
it is interesting to note that Pecina and Ceplecha (1983, 1984) state these
equations are "violated" for longer trajectories, indicating the simplicity of
their chosen velocity models.
1.2. Straight Line Least Squares (Borovicka, 1990)
Only three years following MOP, the Straight Line Least Squares (SLLS)
method was published. Although Borovicka (1990) showed the SLLS method
to produce lower residuals than MOP, they concluded that both methods
produce similar results and could not recommend one over the other; even
suggesting a combination of both may be preferable, depending on the case.
That said, Gural (2012) found the SLLS method to be more robust when
lower resolution cameras were used.
3
Unlike MOP, the SLLS method best fits a straight line trajectory di-
rectly to all the observed lines-of-sight at once.
It does so by minimising
the perpendicular distances between the lines-of-sight and the straight line
trajectory itself. It was later stated by Gural (2012) that a better alternative
to the initially published SLLS method was to minimise the angular distance
rather than the perpendicular distance. Using the angular distance acts to
indirectly weight the line-of-sight measurements based on their observation
range.
The positions are determined for every line-of-sight by determining the
closest point on the optimised straight line trajectory to that given line-of-
sight (regardless of time). Similar to MOP, the SLLS method requires a
separate step to determine the velocity along the trajectory. The methods
of Pecina and Ceplecha (1983, 1984) are used to determine this velocity by
considering the 1D lengths along the trajectory over time.
We must note that Borovicka (1990) offers the SLLS method in both the
Earth Centered/Earth Fixed (ECEF) frame and the Earth Centred Inertial
(ECI) frame; the main difference is where the straight line trajectory is de-
fined. Performing the SLLS method in the ECI frame implicitly includes the
Coriolis force, but requires absolute timing knowledge to operate. It is up to
the user to determine which variation is more physically realistic.
1.3. Multi-Parameter Fit (Gural, 2012)
The previously discussed MOP and SLLS methods are purely geometric
triangulation solutions; i.e.
the trajectory fitting component can be per-
formed without any timing information. It is only as a second step, when
velocity analysis is needed, that timing of the observed meteoroid is con-
sidered. This means MOP and SLLS determine a unique position for every
line-of-sight; if there are simultaneous observations from N sensors, there will
be N unique positions along the trajectory corresponding to the same point
in time. Only later, in the velocity analysis step, can this potential scatter
be dealt with.
The Multi-Parameter Fit (MPF) technique of Gural (2012) differs from
the previous two triangulation methods in that it fits raw observations di-
rectly to a trajectory solution, combining the straight line fitting and velocity
modelling steps into one. Hence, N simultaneous observations will now result
in one unique position along the trajectory. One implication of this approach
is that the convergence angle can now be thought of as the angle between
simultaneous lines-of-sight rather than between planes, which is a significant
distinction.
As the name suggests, the MPF algorithm best fits unknown trajectory
parameters to the measured lines-of-sight by minimising the angular dis-
4
tance between said lines-of-sight and the predicted lines-of-sight given their
modelled positions along a straight line trajectory. These fitting parameters
include the initial position ((cid:126)p0), the initial velocity ((cid:126)v0), some deceleration
coefficients depending on the chosen model (ai), and sensor timing offsets
(∆tk), giving the MPF the ability to handle asynchronous sensors assuming
they all have relative timing. Positions along the straight line trajectory are
determined using one of three velocity models: a constant velocity along the
track, a linearly decreasing velocity with time, or an exponentially dependent
deceleration (Jacchia et al., 1967). However, these suggested velocity models
do not physically represent the trajectory dynamics. Gural (2012) suggests
that this technique is most applicable to smaller mass meteors (< 5 g) of
short duration (< 3 sec), unless a better model is used.
2. New Approach - Dynamic Trajectory Fit
Of the three most prevalent triangulation methods (as discussed in Sec-
tion 1), none claim to be able to fit long-duration fireballs of significant mass,
in part because all methods have assumed a straight line trajectory. While
Jenniskens (2006) claim that masses < 50 g or equivalent magnitude of -2
can be approximated using straight line trajectories, if the goal is to observe
deeply penetrating fireballs such as those targeted by the GFO, the fireballs
are not guaranteed to follow this straight line assumption. In fact, Sansom
et al. (2019) show that the straight line assumption is an oversimplification
that will affect orbit calculations and meteorite search regions for a significant
number of fireball events.
The Dynamic Trajectory Fit (DTF) method proposed here removes this
straight line assumption by fitting differential equations of motion directly
to the measured lines-of-sight, thereby including all spacial/temporal infor-
mation in one step and ultimately providing a more realistic account of the
meteoroid's fall trajectory. This methodology takes the ideas of global fitting
proposed in Gural (2012) several steps further. The differential equations
that describe meteoroid fall dynamics and ablation are as follows (Sansom
et al., 2015):
(cid:13)(cid:13)(cid:126)vrel
(cid:13)(cid:13)
d(cid:126)v
dt
dm
dt
(cid:126)vrel + (cid:126)agrav
= −cdAρa
2ρ2/3
m m1/3
= −chAρam2/3(cid:13)(cid:13)(cid:126)vrel
(cid:13)(cid:13)3
2H∗ρ2/3
m
(1)
(2)
where (cid:126)v is the meteoroid's absolute velocity in the Earth-Centred Inertial
(ECI) frame, (cid:126)vrel is the meteoroid's velocity relative to the atmosphere, m is
5
the meteoroid's mass, (cid:126)agrav is the acceleration due to gravity1, cd is the drag
coefficient, A is the shape-density parameter (Bronshten, 1983), ρm is the
meteoroid's density, ρa is the atmospheric air density2, ch is the heat-transfer
coefficient, and H∗ is the enthalpy of sublimation.
However, not every unknown parameter from Eq. 1 and Eq. 2 can be re-
solved as many terms are dynamically coupled, and therefore indistinguish-
able given only the line-of-sight measurements we obtain. Therefore, we can
alter these equations by grouping the coupled terms together as shown:
= −ρa
d(cid:126)v
dt
(cid:126)vrel + (cid:126)agrav
(cid:13)(cid:13)(cid:126)vrel
2β
(cid:13)(cid:13)
(cid:13)(cid:13)(cid:126)vrel
6
(cid:13)(cid:13)3
= −σρa
dβ
dt
(3)
(4)
where β = 3(cid:112)mρ2
m/(cdA) = m/(cdS) is the meteoroid's ballistic coeffi-
cient, σ = ch/(cdH∗) is the meteoroid's ablation coefficient, (cid:126)vrel = (cid:126)v − (cid:126)vatm
is the meteoroid's velocity relative to the atmosphere, (cid:126)vatm = (cid:126)ωe × (cid:126)p is the
velocity of the atmosphere, (cid:126)ωe is Earth's rotational angular velocity, (cid:126)p is
the meteoroid's position in the ECI frame, and S is the meteoroid's cross-
sectional area.
In addition to estimating the dynamic parameters ((cid:126)p and (cid:126)v), by fitting the
above differential equations to the measurements, the DTF method can also
estimate some physical parameters, including the ballistic parameter, β, and
ablation coefficient, σ. By assuming a constant value of the meteoroid's shape
and density throughout its luminous trajectory, the fitted ballistic parameter,
β, can be used to estimate the meteoroid's mass during its observed decent
through the atmosphere - see Section 2.1 for details.
Although some fireball networks have sub-millisecond timing precision on
their shutter actuations within a long-exposure image, such as those obser-
vatories within the GFO (Howie et al., 2017b), the identification of the exact
shutter breaks is not as precise due to halo-ing and/or saturation of the fire-
ball. To determine any missing temporal information due to these effects,
provided at least one sensor has timing for reference, the DTF method is
able to handle observations without timing all-together. Additionally, the
DTF method can resolve for any timing offsets between sensors, which is
1Care must be taken in calculating the direction of the Earth's gravitation vector. It
should be perpendicular to Earth's ellipsoid rather than towards Earth's center-of-mass;
a subtle, but accumulative difference.
2The atmospheric density, ρa, is calculated using the NRLMSISE-00 empirical atmo-
spheric model (Picone et al., 2002).
6
necessary for those meteor and fireball networks that only record relative
time information.
One extra feature available as a consequence of the DTF approach is
the option to include fragmentation events, which can be user-diagnosed by
large flares in the light-curve.
If prompted, the DTF method can resolve
for both time and amount of discrete fragmentation using the deceleration
characteristics of the meteoroid inherent in the observations.
2.1. Procedure
A conceptual overview of the Dynamic Trajectory Fit (DTF) methodol-
ogy will be presented here, with sufficient detail to ensure reproducibility. For
reference and/or use, the Python source code will be made publicly available
on the DFN's GitHub3.
Computationally, the DTF algorithm is divided into three main parts:
state approximation, sanity checks, and optimisation.
Part 1: State Approximation. In preparation for the main optimisation step
(Part 3), we must estimate all the unknown parameters for a single point
in time that describe the meteoroid's dynamics, see Eq. 3 and Eq. 4. The
collection of these parameters is termed the meteoroid's "state", and is given
by the following vector:
(cid:126)χest = [pf
x, pf
x, pf
y , pf
z , vf
x, vf
y , vf
y , pf
z ] is the final position, (cid:126)vf = [vf
z , βf , σ, δf rag,i, tf rag,i, ∆tj, trel
k ]
(5)
y , vf
x, vf
where (cid:126)pf = [pf
z ] is the final ve-
locity, βf is the final ballistic coefficient, and σ is the ablation coefficient. We
use the end of the observable trajectory in the state estimate as it is far easier
to constrain the ballistic coefficient, which relates to meteoroid mass, to be
greater than zero for all times along the observable trajectory (β(t) > 0).
These first 8 parameters are always required to define the trajectory. If one
or more fragmentation events are suspected, the percentage fragmentation,
δf rag, and the time of fragmentation, tf rag, are added to the state for ev-
ery possible event. If one or more observatories are found to contain timing
offsets, an estimated offset time, ∆t, is added to the state for every offset ob-
servatory. Finally, if one or more observatories contain lines-of-sight without
relative times, an estimated relative time, trel, is added for every line-of-sight
that is missing timing information.
To calculate these estimates, we must first get an idea of the trajectory
from simpler triangulation methods. Using a boot-strapping approach, we
3Please follow https://github.com/desertfireballnetwork/ for the source code to
the DTF algorithm.
7
can build-up from the MOP (Ceplecha, 1987) to the SLLS (Borovicka, 1990),
from which we can then estimate most of the state parameters. The time
components are estimated first to ensure we calculate the correct position,
velocity, and ballistic coefficient parameters.
To determine if any observatories have a timing offset problem, we start
by assuming all the sensors are offset and determine what ∆t is needed to
synchronise them. This involves first designating a "master" observatory to
act as a temporal anchor, chosen as the observatory with the most lines-of-
sight with timing. Next we adjust the estimated timing offsets for every other
observatory to minimise the differences in lengths along the SLLS line when
compared to the "master", interpolating if necessary. If the estimated offset
is greater than a given tolerance, say 0.05 seconds, then the timing from that
observatory is used as relative timing only, and the estimated offset is added
to the state to be optimised.
All the lines-of-sight without timing are then very roughly estimated by
comparing their lengths along the SLLS line to a modelled length/time func-
tion. This function is constructed by fitting a trajectory of constant velocity
to the SLLS lengths along the line over time. All timeless lines-of-sight have
their along-track lengths converted to relative timing, trel, and are subse-
quently added to the state estimate to be optimised. After optimisation,
these lines-of-sight each produce zero along-track error, as expected.
Now using the rough timing corrections above, we are able to more ac-
curately estimate the meteoroid's final position and velocity from the SLLS
fit. Put simply, the estimated position is merely the final triangulated point
along the SLLS line, and the estimated velocity is a least-squares average
velocity of the last eight SLLS triangulated positions. The ablation coeffi-
cient, σ, is initially estimated as 14 × 10−9 s2/m2 in all cases (Sansom et al.,
2015). The ballistic coefficient, βf , is roughly determined by equating the
SLLS trajectory length with the propagated trajectory length assuming a
βf value. This is achieved using Brent's root-finding method on the range
log10(βf ) ∈ [1, 4], which approximately equates to a meteoroid mass range of
0.1 g to 100 ton sphere of chondritic density (3500 kg/m3).
If any fragmentation is suspected by the user, one or more fragmentation
times are able to be input to the algorithm and serve as the tf rag parameter
in the state estimate. The fragmentation percent, δf rag is always estimated
initially as 30%, and adjusted upon optimisation.
We must note that these estimates' sole purpose is to start the optimi-
sation sufficiently close to the global minimum to allow convergence. Once
the minimisation algorithm begins, the measurements are the only things
directly influencing the trajectory solution; it is not building off already pro-
cessed data.
8
Part 2: Sanity Checks. As some fireball data reduction pipelines can be
almost completely automated, such as that of the GFO, there is a chance
sensor data is corrupt or has been incorrectly grouped. This could occur for a
variety of reasons, including calibration errors, planes and/or satellites being
misidentified as meteoroids, or the rare cases where multiple simultaneous
fireballs are incorrectly correlated across sensors.
To avoid triangulation errors within the optimisation routine, a variety of
sanity checks need to be performed to remove erroneous data before the op-
timisation is attempted. All the following checks use the rough triangulation
of SLLS and the state approximation (as determined in Part 1) to ensure
that each observatory's triangulated observations:
1. Decrease in height over time.
2. Change in height at roughly the same rate.
3. Produce sufficiently low SLLS residuals.
4. Triangulate to positions above the ground.
5. Triangulate to positions less than 200 km altitude.
6. Produce a final state velocity estimate less than 200 km/s.
These conditions are designed to be quite extreme to prevent accidentally
discarding any valid data that has happened to triangulate poorly using the
SLLS procedure. If any data is found inaccurate, the first sensor to fail an
above condition is eliminated and the procedure begins over from the state
approximation (Part 1).
Part 3: Optimisation. Now that we have an initial state estimate (Part 1)
using good data (Part 2), we are now in a position to begin the trajectory
optimisation. This step could be performed with any robust minimisation
routine that imposes bounds on the optimised state to ensure realistic results.
For reliability, we have elected to use SciPy's in-built least-squares function
that has been thoroughly tried and tested (Virtanen et al., 2019). Within
this function, the Trust Region Reflective (TRF) method is chosen as it
is robust and permits bounds to be set on the allowable state. We define
rather generous state bounds to give the optimisation routine enough room to
effectively search the state-space while at the same time keeping the resulting
state physically realistic, see Table 1.
The chosen TRF method also offers the option for user-defined Jacobian
and state step-size. For accuracy and computational speed, we provide a
parallelised custom Jacobian function that uses central differencing. The
step-size is defined equal to the change in state used in the Jacobian's central
differencing algorithm to avoid state divergence by overshooting the bounds
of Jacobian linearity.
9
Table 1: State boundary conditions given to the least-squares algorithm to ensure realistic
results, where LB and UB stand for lower-bound and upper-bound respectively. Also, the
star-symbol represents the associated estimated state parameter as determined in Part 1.
State
Parameters
LB
UB
(cid:126)pf
(km)
* - 40
* + 40
(cid:126)vf
(km/s)
* - 5
* + 5
β
(kg/m2)
10−10
104
σ
(kg/J)
3 × 10−9
3 × 10−6
δf rag
(%)
0
100
tf rag ∆t, trel
(s)
tmin
tmax
* - 10
* + 10
(s)
Once the least-squares algorithm is setup and initiated, the state is prop-
agated using Eq. 3 and Eq. 4 to all the other observation times and subse-
quently converted to lines-of-sight. These predicted lines-of-sight are differ-
enced from the observed lines-of-sight to give the angular along-track and
cross-track residual components. With the help of the Jacobian to show the
direction of the local (and hopefully global) minimum, the state parameters
are adjusted to minimise these angular residuals, weighted by their individual
astrometric uncertainties. This procedure occurs iteratively until the state
does not differ significantly enough from one iteration to the next; therefore
signifying that a minimum is reached and the resulting state matches the
observations as closely as possible.
Now that the optimised state solution is obtained, the state errors are
determined (as discussed in Section 2.2) and propagated to all the other
observation times alongside the state itself before being saved to file for sub-
sequent orbit determination and possible darkflight analysis. Various plots
are then constructed using this data, see Section 3.2.
2.2. Notes on Errors
We must note that the least-squares algorithm used within the DTF
method does not produce errors. Instead, covariance errors can be estimated
afterwards from both the Jacobian of the optimised state and the covariance
on the line-of-sight measurements as follows (Bevington et al., 1993):
(cid:126)χcov = (cid:126)χres
cov + (cid:126)χz
cov
cov = (d(cid:126)χ/d (cid:126)res)T diag( (cid:126)res2)(d(cid:126)χ/d (cid:126)res)
(cid:126)χres
cov = (d(cid:126)χ/d (cid:126)res)T (d (cid:126)res/d(cid:126)z)T (cid:126)zcov(d (cid:126)res/d(cid:126)z)(d(cid:126)χ/d (cid:126)res)
(cid:126)χz
(6)
(7)
(8)
(9)
where (cid:126)J is the state Jacobian matrix, describing how the residuals change
with a change in state; d(cid:126)χ/d (cid:126)res is the inverse of the Jacobian, describing how
d(cid:126)χ/d (cid:126)res = ( (cid:126)J (cid:126)J T )−1 (cid:126)J T
10
the state changes with a change in residuals; d (cid:126)res/d(cid:126)z is a coordinate trans-
form, describing how the residuals (along-track/cross-track) change with a
change in line-of-sight measurements (ra/dec); (cid:126)zcov is the covariance on the
measurements; and diag( (cid:126)res2) is the residual vector at the optimised state,
diagonalised.
As shown in Eq. 6, we are able to incorporate the residual covariance
due to the spread in residuals around the model, (cid:126)χres
cov, and measurement
covariance due to the astrometric uncertainty, (cid:126)χz
cov, into an overall covari-
ance estimate. Separate testing showed that the measurement covariance
component accurately reflected the covariance of the state through repeated
Monte-Carlo analyses in which the measurements were varied within their
astrometric covariance space.
However, we must also make note that the uncertainty formulation dis-
cussed above does not account for errors arising due to the meteoroid equa-
tions of motion as well as assumptions made within this model (Eq. 3 and
Eq. 4), such as a constant ablation coefficient, shape, and density of the mete-
oroid throughout the visible trajectory. Therefore, the determined covariance
from Eq. 6 can be viewed as minimum uncertainties given the observations.
3. Results and Discussion
To demonstrate and compare the capabilities of the four previously dis-
cussed triangulation methods, we conduct two independent comparative
analyses: The first study uses over 100,000 randomly simulated trajectories,
comparing the fitted initial velocity vector to the simulated "truth". The
second study uses a real fireball event, captured by multiple observatories
within the Desert Fireball Network.
3.1. Randomised Simulations
To fully analyse the accuracy of a triangulation algorithm through the
full range of possible trajectory conditions, one must rely on simulation.
Simulations allow us to compare a triangulation solution against the unal-
tered trajectory "truth". For the following comparative analysis, a fireball
simulator was designed, built, and heavily tested under a variety of initial
conditions before being used to compare the various triangulation methods.
This fireball simulator begins with a set of randomised physical and dy-
namical initial conditions at the top of the atmosphere, that completely de-
fines a meteoroid's state at that point. This randomised state is then nu-
merically propagated forward in time using the meteoroid's 3D differential
equations of motion until the meteoroid's speed relative to the ground falls
11
below 2 km/s. Likewise, the initial meteoroid's state is also propagated back
in time until the meteoroid's height exceeds 200 km.
Once this simulated trajectory has been established, perfect azimuth and
elevation measurements are generated every 0.1 seconds for two (or more)
randomised observatory locations for the section of the trajectory that would
be visible to the sensor - i.e. while the meteoroid is more than 10◦ above
the horizon and ablating rapidly enough to be detectable from each obser-
vatory's perspective. These resulting measurements are then varied within
some randomised Gaussian measurement error to better reflect reality4.
The initial state of these simulated trajectories was generated with a fixed
latitude of 0◦, a fixed longitude of 0◦, a fixed height of 100 km, a uniformly
random slope between 10◦ and 90◦, a uniformly random bearing between 0◦
and 360◦, and a uniformly random speed between 12 km/s and 72 km/s. Ad-
ditionally, the meteoroid was initialised with a fixed density of 3500 kg/m3,
a fixed spherical shape, and a uniformly random mass (in log-space) between
100 g and 100 kg. Two uniformly random observatory locations were gener-
ated that could view the centre of the observable trajectory at an elevation
greater than 20◦. This did not always generate geometrically favourable ob-
servation combinations.
The simulated line-of-sight observations were given measurement error
of 2.4 arcmin, characteristic of the measurement errors given by a Desert
Fireball Network observatory (Howie et al., 2017a). Gural (2012) found
that the resulting radiant error was proportional to measurement error, and
therefore any results found through this analysis can be linearly extrapolated
to imaging systems of higher or lower resolution.
In this analysis, we generated 123,337 sets of realistic double-station mea-
surements from random trajectories using the fireball simulator. Each mea-
surement set was subsequently passed to the four triangulation methods for
trajectory fitting: the Method of Planes (MOP), the Straight Line Least
Squares (SLLS) method, the Multi-Parameter Fit5 (MPF) method, and the
novel Dynamic Trajectory Fit (DTF) method. The original simulated radi-
ant velocity vector and the four fitted radiant velocity vectors from the top
of the trajectory are then compared, distinguishing the differences in slope,
4This trajectory can also be effected by multiple randomised or user defined fragmen-
tation events, and/or systematic observatory timing offsets to increase realism, however
these abilities are not used in this analysis.
5Gural (2012) states that "the algorithm is not ill-conditioned to having too many
velocity velocity fitting parameters as long as there is measurement sample support."
Therefore, we have chosen to use the exponentially dependent deceleration model specified
in Eq. 4 of Gural (2012) for MPF analysis within this paper.
12
bearing, and velocity magnitude components.
Similar to the analysis performed by Gural (2012), the difference between
the true and estimated radiant parameters are statistically analysed by con-
sidering its median value within small, equally divided bins that subtend the
x-axis. This avoids excess clutter and highlights the general trends of the
various triangulation methods.
Using the approach described above, we can compare the fitting errors
against different meteoroid trajectory parameters, such as observation con-
vergence angle, initial speed, trajectory duration, and trajectory length as
shown in Fig. 1, Fig. 2, Fig. 3, and Fig. 4, respectively.
Figure 1: The median absolute differ-
ences between the simulated and the fit-
ted radiant at the top of the meteoroid's
trajectory by varying observation con-
vergence angle.
Figure 2: The median absolute differ-
ences between the simulated and the fit-
ted radiant at the top of the meteoroid's
trajectory by varying initial speed.
13
Figure 3: The median absolute differ-
ences between the simulated and the fit-
ted radiant at the top of the meteoroid's
trajectory by varying trajectory dura-
tion, where duration is roughly propor-
tional to number of collected observa-
tions.
Figure 4: The median absolute differ-
ences between the simulated and the fit-
ted radiant at the top of the meteoroid's
trajectory by varying trajectory length,
where length can be roughly related to
initial speed and trajectory duration.
From these simulation results, we notice that all triangulation methods
generally agree and tend to follow the same trends. Areas of most model
inaccuracy arise when a meteoroid trajectory is viewed from observatories
of low convergence angle, is short in length, or displays a relatively slow
entry velocity.
Interestingly, these are the regions that the DTF method
either matches or exceeds in accuracy when compared to the alternative
triangulation methods.
In particular, the DTF provides a more accurate
trajectory solution at low convergence angles (< 10◦), slow to moderate entry
velocities (< 25 km/s), and extremely fast entry velocities (> 65 km/s).
Regions where the DTF method appears to perform poorly could be due
to the underlying least-squares algorithm either reaching a non-global min-
imum or simply terminating optimisation procedures too early. Regardless,
the estimated errors calculated as part of the DTF procedure (Section 2.2)
are on the same order as the median absolute deviations shown from these
simulations. This indicates that the true meteoroid trajectory is accurately
encompassed within the DTF errors, which is the ultimate goal of meteoroid
14
trajectory modelling.
It is also interesting to note that in most trajectory scenarios, the mod-
elled velocity error is on the order of 0.1 km/s. However, as stated before
in Gural (2012), the magnitude of this model error is directly proportional
to the uncertainty in the line-of-sight observations. Therefore, we can con-
clude that meteoroid events with observation errors less than the 2.4 arcmin
simulated here should result in a velocity accuracy better than ∼0.1 km/s -
the threshold needed for accurate identification of meteoroid source regions
within the Solar System (Granvik and Brown, 2018).
3.2. Case Study: Fragmentation Event (DN141125 01)
Simulations are a way to thoroughly investigate and compare various
models to an estimated reality. However, no simulation can 100% replicate
reality. It is for this reason that we analyse and compare the various mete-
oroid triangulation methods using a real-world example. We choose an event
with visible signs of fragmentation to highlight the fragmentation handling
within the DTF method, as shown in Fig. 5.
Figure 5: The captured long-exposure image of event DN141125 01 taken from the Mul-
gathing station within the Desert Fireball Network, showing visible signs of fragmentation
towards the end of the luminous trajectory.
This fireball event with visible fragmentation, referred to as DN141125 01,
was captured by five DFN observatories - two of which could not be resolved
for timing due to the distance of the observations. Although the DTF method
can incorporate data with this lack of timing information, we chose to discard
the data from these observatories for triangulation comparison purposes. The
DN14125 01 event was visible for 9.24 seconds, comprising of 459 line-of-sight
observations at a maximum convergence angle of 35◦. The triangulation for
event DN141125 01 is shown visually in Fig. 6 and is summarised in Table 2.
15
Figure 6: The triangulation of event DN141125 01 using a total of 459 line-of-sight mea-
surements from three South Australian observatories within the Desert Fireball Network:
Mulgathing, Northwell, and Mount Ives.
To determine which triangulation model best fits the line-of-sight obser-
vations, we compare the residual magnitudes as stated in Table 2 and shown
more thoroughly in Fig. 7. Unsurprisingly, the residuals in the cross-track
direction are smallest using the SLLS method as this is its optimisation pa-
rameter. However, the DTFf rag model possesses the smallest total residuals.
16
Table 2: Summary of event DN141125 01 trajectory parameters using the four triangu-
lation methods discussed in this paper: the Method of Planes (MOP), the Straight Line
Least Squares (SLLS), the Multi-Parameter Fit (MPF), and the Dynamic Trajectory Fit
(DTF). In addition, the triangulation solution of the Dynamic Trajectory Fit with frag-
mentation (DTFf rag) was also given to highlight this added fitting feature. The results are
divided into four sections; the standard deviations of the trajectory residuals to indicate
goodness of fit, the radiant direction for possible meteor stream classification, the initial
trajectory position and velocity at 15:21:15.386 UTC used for orbit determination, and the
final trajectory position and velocity at 15:21:24.626 UTC used for darkflight analysis.
ATR [']
CTR [']
Total [']
RA∞ [◦]
Dec∞ [◦]
Lat0 [◦]
Lon0 [◦]
Hei0 [km]
Vel0 [ km
s ]
Slope0 [◦]
Azi0 [◦]
Mass0 [kg]
Latf [◦]
Lonf [◦]
Heif [km]
Velf [ km
s ]
Slopef [◦]
Azif [◦]
Massf [kg]
MOP
8.700
2.427
9.033
345.257
-46.398
-31.593
133.770
80.441
13.977
26.710
230.046
N/A
-31.011
134.545
30.456
4.711
26.705
230.043
N/A
SLLS MPF
6.385
4.099
0.861
2.314
6.792
4.188
345.088
345.101
-46.663
-46.701
-31.600
-31.600
133.765
133.767
80.815
80.752
14.095
14.381
27.236
26.712
229.688
228.656
N/A
-31.011
134.541
30.627
4.954
26.707
229.686
N/A
N/A
-31.012
134.538
30.732
3.041
27.232
228.654
N/A
DTF DTFf rag
2.543
3.392
4.240
345.014
-46.333
-31.593
133.768
80.285
13.989
26.532
230.054
0.901
-31.010
134.539
30.543
4.738
25.822
231.889
0.081
2.332
3.411
4.132
345.010
-46.311
-31.592
133.769
80.189
13.908
26.520
230.075
1.605
-31.010
134.540
30.521
4.892
25.861
231.803
0.113
17
Figure 7: The along-track and cross-track residuals from all three observatories of the
DN141125 01 event using five triangulation methods: the four discussed methods as well
as the Dynamic Trajectory Fit method with fitted fragmentation (DTFf rag). The standard
deviation of these residuals are given in Table 2.
The velocities determined by the various triangulation methods rely on
different models, each containing unique assumptions. The velocity deter-
mination algorithm used within the MOP and SLLS methods fits the 1D
meteoroid equations of motion to the lengths along the 1D trajectory, as-
suming an exponential atmosphere Pecina and Ceplecha (1983). The veloc-
ity calculated by the MPF method uses a purely empirical formula (Whipple
and Jacchia, 1957; Gural, 2012). Lastly, the velocity results from the DTF
method consults the meteoroid's 3D equations of motion directly, without any
simplifying straight line or atmospheric assumptions. The subtleties between
these velocity models using data from event DN141125 01 are compared in
Fig. 8.
18
Figure 8: The modelled velocity of the DN142511 01 event from the various triangula-
tion methods, as discussed in Section 1 and Section 2. The surrounding scatter is the
instantaneous velocities as calculated by the change in adjacent SLLS positions along the
trajectory over the change in time from each observatory, separately. Velocity subtleties
at the beginning, middle, and end of the trajectory are highlighted by zoomed in sections.
As shown in Table 2 and Fig. 8, the final velocity predicted by the MPF
method does not appear to follow the instantaneous velocity scatter, sug-
gesting the exponentially dependent velocity model does not reflect reality
for long fireball-type events. Excluding the MPF velocity, the remaining
velocity models seem very similar, varying by about 300 m/s at the extremi-
ties. However, this 300 m/s variation would still lead to considerably different
darkflight and orbit regression results.
As discussed in Section 2, the DTF method is able to resolve for the
meteoroid's ballistic coefficient over time, β(t). By assuming a constant me-
teoroid shape and density, we can estimate the meteoroid's mass throughout
the observed luminous trajectory directly using the line-of-sight observations
- unlike any other compared triangulation method. This feature not only
helps diagnose meteorite-dropping events, but assists greatly in constraining
the meteorite search area. The mass estimates for event DN141125 01 using
the DTF and DTFf rag methods are compared in Fig. 9. The DTFf rag method
predicts the meteoroid from DN141125 01 broke up around 5.3 seconds, at
an altitude of 47.3 km - consistent with the visible fragmentation shown in
Fig. 5.
19
Figure 9: The estimated mass of the DN141125 01 event throughout its trajectory deter-
mined by the Dynamic Trajectory Fit, both with (DTF) and without (DTFf rag) fragmen-
tation fitting. The other three triangulation models are not plotted here as they do not
produce mass estimates.
To summarise this comparison of triangulation methods, the Dynamic
Trajectory Fit with fragmentation handling (DTFf rag) appears to be the
best model for event DN141125 01. While there may be events that are
better suited to the other triangulation methods, the simulations discussed
in Section 3.1 show that the DTF method is an equal if not better choice
for most events. Additionally, the DTF method can estimate the mass of
the meteoroid from the line-of-sight observations directly, as discussed in
Section 1.1, Section 1.2, and Section 1.3.
4. Future Functionality
While the proposed DTF method appears successful in its current form,
and poses considerable merit, there are a few improvements that could be
applied to increase realism and draw out additional subtleties within the
gathered data. These improvements include:
Light-curve incorporation. With the inclusion of light-curve data, we would
have the opportunity to better model meteoroid mass-loss along the tra-
jectory, which would act to further constrain the meteoroid state and its
associated uncertainty. Luminous efficiency models, such as Gritsevich and
20
Koschny (2011), could be relatively easily incorporated into the state prop-
agation of the meteoroid to better estimate its physical and dynamical pa-
rameters.
Automated fragmentation determination. Currently, we rely on a user-
defined time of fragmentation. However, with full light-curve history, we
should be able to flag fragmentation events from light-curve peaks alone,
therefore negating any user-required input to the algorithm. However, this
functionality could easily be integrated upstream in a larger data reduc-
tion pipeline using measurements from highly sensitive radiometers (Buchan
et al., 2019), not necessarily integrated in the triangulation method itself.
Meteoroid spin modelling. For some particularly long fireballs, such as Case 1
of Sansom et al. (2019), trajectories appear to considerably deviate from the
fall-plane, suggesting there are unaccounted aerodynamic effects. We hy-
pothesise this might be in part due to the Magnus Effect at high velocities;
that is, the resulting curvature of an objects trajectory due to its spin. It
would be very interesting to model these cases with meteoroid spin consid-
ered. The proposed DTF method would simply require an additional three
state parameters to model this phenomena; namely the angular velocity vec-
tor, (cid:126)ωspin = [ωx, ωy, ωz]. It would be conceivable to extend the DTF method
to optimise without spin, only re-optimising with spin if the measurements
did not adequately match the model (reduced chi squared χ2
ν ≈ 1).
5. Conclusions
Meteoroid orbits and meteorite samples provide invaluable information
that helps planetary scientists investigate Solar System formation and the
origin of life on Earth. Fireball networks around the globe are on the forefront
of providing this knowledge. However, the accuracy of the determined orbit
and the chance of meteorite recovery both rely heavily on the accuracy of
the underlying meteoroid triangulation method.
Three triangulation methods have been proposed in the past:
the
Method of Planes (Ceplecha, 1987), the Straight Line Least Squares method
(Borovicka, 1990), and the Multi-Parameter Fit (Gural, 2012). The first two
listed methods above separate out the geometric fit from the dynamic mod-
elling. In 2012, Gural simplified this procedure to a single step, changing
the well-known convergence angle from that between planes to that between
simultaneous rays - a clear advantage over the past traditional triangulation
methods. However, the velocity models suggested within Gural (2012) are
empirically derived for small meteors and do not reflect reality, particularly
for meteorite-dropping events. The proposed novel Dynamic Trajectory Fit
21
method not only contains a more realistic dynamic model, but it possesses the
ability to determine the meteoroid's ballistic coefficient throughout the ob-
servable trajectory directly from the line-of-sight measurements - unlike any
other proposed triangulation method. With meteoroid shape and density
assumptions, this ballistic coefficient can be easily translated into meteoroid
mass.
Over 100,000 multi-station meteoroid simulations revealed the advantage
of the Dynamic Trajectory Fit method particularly for relatively slow entry
events (<25 km/s) as well as events observed from low convergence angles
(<10◦). Additionally, a visibly fragmenting fireball event captured by three
stations of the Desert Fireball Network was used to compare the four tri-
angulation methods. The Dynamic Trajectory Fit with fragmentation was
shown to best match the observations, with the predicted fragmentation time
in agreement with the observed data.
The method proposed here could be easily modified to fit arbitrarily com-
plex equations of motion, to include light-curve data, and to provide auto-
mated fragmentation detection in the future.
6. Acknowledgements
This work was funded by the Australian Research Council as part of the
Australian Discovery Project scheme, and supported by resources provided
by the Pawsey Supercomputing Centre with funding from the Australian
Government and the Government of Western Australia. This work was also
supported by an Australian Government Research Training Program (RTP)
Scholarship.
This research made use of Astropy, a community-developed core Python
package for Astronomy (Collaboration et al., 2013). Additionally, the major-
ity of figures were generated using Matplotlib, another community-developed
Python package (Hunter, 2007).
7. References
References
P. R. Bevington, D. K. Robinson, J. M. Blair, A. J. Mallinckrodt, and
S. McKay. Data Reduction and Error Analysis for the Physical Sci-
ences. Computers in Physics, 7:415, 1993.
ISSN 0894-1866. URL
http://adsabs.harvard.edu/abs/1993ComPh...7..415B.
J. Borovicka. The comparison of two methods of determining meteor tra-
jectories from photographs. Bulletin of the Astronomical Institutes of
22
Czechoslovakia, 41:391 -- 396, Dec. 1990.
//adsabs.harvard.edu/abs/1990BAICz..41..391B.
ISSN 0004-6248. URL http:
V. A. Bronshten. Physics of meteoric phenomena. 1983. URL http://
adsabs.harvard.edu/abs/1983pmp..book.....B.
S. R. G. Buchan, R. M. Howie, J. Paxman, and H. A. R. Deville-
poix. Developing a Cost-Effective Radiometer for Fireball Light Curves.
arXiv:1907.12807 [astro-ph], July 2019. URL http://arxiv.org/abs/
1907.12807. arXiv: 1907.12807.
Z. Ceplecha. Multiple fall of Pibram meteorites photographed. 1. Double-
station photographs of the fireball and their relations to the found me-
teorites. Bulletin of the Astronomical Institutes of Czechoslovakia, 12:
21, 1961. URL https://ui.adsabs.harvard.edu/abs/1961BAICz..12.
..21C/abstract.
Z. Ceplecha. Geometric, dynamic, orbital and photometric data on mete-
oroids from photographic fireball networks. Bulletin of the Astronomical
Institutes of Czechoslovakia, 38:222 -- 234, July 1987. ISSN 0004-6248. URL
http://adsabs.harvard.edu/abs/1987BAICz..38..222C.
T. A. Collaboration, T. P. Robitaille, E. J. Tollerud, P. Greenfield, M. Droet-
tboom, E. Bray, T. Aldcroft, M. Davis, A. Ginsburg, A. M. Price-Whelan,
W. E. Kerzendorf, A. Conley, N. Crighton, K. Barbary, D. Muna, H. Fer-
guson, F. Grollier, M. M. Parikh, P. H. Nair, H. M. Gnther, C. Deil,
J. Woillez, S. Conseil, R. Kramer, J. E. H. Turner, L. Singer, R. Fox,
B. A. Weaver, V. Zabalza, Z. I. Edwards, K. A. Bostroem, D. J. Burke,
A. R. Casey, S. M. Crawford, N. Dencheva, J. Ely, T. Jenness, K. Labrie,
P. L. Lim, F. Pierfederici, A. Pontzen, A. Ptak, B. Refsdal, M. Servil-
lat, and O. Streicher. Astropy: A Community Python Package for As-
tronomy. July 2013. doi: 10.1051/0004-6361/201322068. URL https:
//arxiv.org/abs/1307.6212.
M. Granvik and P. Brown. Identification of meteorite source regions in the
Solar System.
ISSN 00191035. doi:
10.1016/j.icarus.2018.04.012. URL http://arxiv.org/abs/1804.07229.
arXiv: 1804.07229.
Icarus, 311:271 -- 287, Sept. 2018.
M. Gritsevich and D. Koschny. Constraining the luminous efficiency of me-
teors. Icarus, 212(2):877 -- 884, Apr. 2011. ISSN 0019-1035. doi: 10.1016/
j.icarus.2011.01.033. URL http://www.sciencedirect.com/science/
article/pii/S0019103511000443.
23
P. S. Gural. A new method of meteor trajectory determination applied to
multiple unsynchronized video cameras. Meteoritics & Planetary Science,
47(9):1405 -- 1418, Sept. 2012. ISSN 1945-5100. doi: 10.1111/j.1945-5100.
2012.01402.x. URL http://onlinelibrary.wiley.com/doi/10.1111/j.
1945-5100.2012.01402.x/abstract.
R. M. Howie, J. Paxman, P. A. Bland, M. C. Towner, M. Cupak, E. K.
Sansom, and H. A. R. Devillepoix. How to build a continental scale fireball
camera network. Experimental Astronomy, 43(3):237 -- 266, June 2017a.
ISSN 1572-9508. doi: 10.1007/s10686-017-9532-7. URL https://doi.
org/10.1007/s10686-017-9532-7.
R. M. Howie, J. Paxman, P. A. Bland, M. C. Towner, E. K. Sansom, and
H. A. R. Devillepoix. Submillisecond fireball timing using de Bruijn time-
codes. Meteoritics & Planetary Science, 52(8):1669 -- 1682, 2017b.
ISSN
1945-5100. doi: 10.1111/maps.12878. URL https://onlinelibrary.
wiley.com/doi/abs/10.1111/maps.12878.
J. D. Hunter. Matplotlib: A 2d Graphics Environment. Computing in Science
Engineering, 9(3):90 -- 95, May 2007. ISSN 1521-9615. doi: 10.1109/MCSE.
2007.55.
L. Jacchia, F. Verniani, and R. E. Briggs. An Analysis of the Atmospheric
Trajectories of 413 Precisely Reduced Photographic Meteors. Smithso-
nian Contributions to Astrophysics, 10:1, 1967. URL https://ui.adsabs.
harvard.edu/abs/1967SCoA...10....1J/abstract.
P. Jenniskens. Meteor Showers and Their Parent Comets. Cambridge Uni-
versity Press, Sept. 2006. ISBN 978-0-521-85349-1. Google-Books-ID: Qpa-
jMuyXG8AC.
P. Pecina and Z. Ceplecha. New aspects in single-body meteor physics.
Bulletin of the Astronomical Institutes of Czechoslovakia, 34:102 -- 121,
Mar. 1983.
ISSN 0004-6248. URL http://adsabs.harvard.edu/abs/
1983BAICz..34..102P.
P. Pecina and Z. Ceplecha. Importance of atmospheric models for interpre-
tation of photographic fireball data. Bulletin of the Astronomical Insti-
tutes of Czechoslovakia, 35:120 -- 123, Mar. 1984.
ISSN 0004-6248. URL
http://adsabs.harvard.edu/abs/1984BAICz..35..120P.
J. M. Picone, A. E. Hedin, D. P. Drob, and A. C. Aikin. NRLMSISE-00
empirical model of the atmosphere: Statistical comparisons and scientific
24
issues. Journal of Geophysical Research: Space Physics, 107(A12):1468,
Dec. 2002.
ISSN 2156-2202. doi: 10.1029/2002JA009430. URL http:
//onlinelibrary.wiley.com/doi/10.1029/2002JA009430/abstract.
E. K. Sansom, P. Bland, J. Paxman, and M. Towner. A novel approach
to fireball modeling: The observable and the calculated. Meteoritics &
Planetary Science, 50(8):1423 -- 1435, Aug. 2015.
ISSN 1945-5100. doi:
10.1111/maps.12478. URL http://onlinelibrary.wiley.com/doi/10.
1111/maps.12478/abstract.
E. K. Sansom, T. Jansen-Sturgeon, M. G. Rutten, P. A. Bland, H. A. R.
Devillepoix, R. M. Howie, M. A. Cox, M. C. Towner, M. Cupak, and
B. A. D. Hartig. 3d Meteoroid Trajectories. Icarus, 321:388 -- 406, Mar.
2019.
ISSN 00191035. doi: 10.1016/j.icarus.2018.09.026. URL http:
//arxiv.org/abs/1802.02697. arXiv: 1802.02697.
P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cour-
napeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der
Walt, M. Brett, J. Wilson, K. J. Millman, N. Mayorov, A. R. J. Nelson,
E. Jones, R. Kern, E. Larson, C. J. Carey, . Polat, Y. Feng, E. W. Moore,
J. VanderPlas, D. Laxalde, J. Perktold, R. Cimrman, I. Henriksen, E. A.
Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa,
P. van Mulbregt, and S. . . Contributors. SciPy 1.0 -- Fundamental Al-
gorithms for Scientific Computing in Python. arXiv:1907.10121 [physics],
July 2019. URL http://arxiv.org/abs/1907.10121. arXiv: 1907.10121.
F. L. Whipple and L. G. Jacchia.
Reduction Methods for Photo-
graphic Meteor Trails. Smithsonian Contributions to Astrophysics, 1:
183, 1957. URL https://ui.adsabs.harvard.edu/abs/1957SCoA....1.
.183W/abstract.
25
|
1810.04961 | 1 | 1810 | 2018-10-11T11:39:19 | Evidence For A Vertical Dependence on the Pressure Structure in AS 209 | [
"astro-ph.EP"
] | We present an improved method to measure the rotation curves for disks with non-axisymmetric brightness profiles initially published in Teague et al. (2018a). Application of this method to the well studied AS$~$209 system shows substantial deviations from Keplerian rotation of up to $\pm 5\%$. These deviations are most likely due to perturbations in the gas pressure profile, including a perturbation located at $\approx 250~$au and spanning up to $\approx 50~$au which is only detected kinematically. Modelling the required temperature and density profiles required to recover the observed rotation curve we demonstrate that the rings observed in $\mu$m scattered light are coincident with the pressure maxima, and are radially offset from the rings observed in mm continuum emission. This suggests that if rings in the NIR are due to sub-$\mu$m grains trapped in pressure maxima that there is a vertical dependence on the radius of the pressure minima. | astro-ph.EP | astro-ph | Draft version October 12, 2018
Typeset using LATEX twocolumn style in AASTeX62
8
1
0
2
t
c
O
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
6
9
4
0
.
0
1
8
1
:
v
i
X
r
a
Evidence For A Vertical Dependence on the Pressure Structure in AS 209
Richard Teague,1 Jaehan Bae,2 Tilman Birnstiel,3 and Edwin A. Bergin1
1Department of Astronomy, University of Michigan, 311 West Hall, 1085 S. University Ave, Ann Arbor, MI 48109, USA
2Department of Terrestrial Magnetism, Carnegie Institution for Science, 5241 Broad Branch Road, NW, Washington, DC 20015, USA
3University Observatory, Faculty of Physics, Ludwig-Maximilians-Universitat Munchen, Scheinerstr. 1, D-81679 Munich, Germany
(Received -; Revised -; Accepted -)
Submitted to ApJ
ABSTRACT
We present an improved method to measure the rotation curves for disks with non-axisymmetric
brightness profiles initially published in Teague et al. (2018a). Application of this method to the well
studied AS 209 system shows substantial deviations from Keplerian rotation of up to ±5%. These
deviations are most likely due to perturbations in the gas pressure profile, including a perturbation
located at ≈ 250 au and spanning up to ≈ 50 au which is only detected kinematically. Modelling
the required temperature and density profiles required to recover the observed rotation curve we
demonstrate that the rings observed in µm scattered light are coincident with the pressure maxima,
and are radially offset from the rings observed in mm continuum emission. This suggests that if rings
in the NIR are due to sub-µm grains trapped in pressure maxima that there is a vertical dependence
on the radius of the pressure minima.
Keywords: editorials, notices -- miscellaneous -- catalogs -- surveys
1. INTRODUCTION
Long baseline observations with the Atacama Large
(sub-)Millimetre Array (ALMA) have shown that sub-
structures in the thermal continuum of protoplanetary
disks are likely ubiquitous. These features are frequently
interpreted in the context of gas pressure maxima into
which grains are shepherded through complex gas-grain
interactions (Birnstiel et al. 2012; Pinilla et al. 2012).
Identification of the main driver of these pressure max-
ima, such as an unseen planet or (magneto-) hydrody-
namical instabilities, is hampered by the lack of a re-
liable tracer of the gas pressure profile. Although CO
isotopologues are routinely found to exhibit structure in
their emission profiles (Isella et al. 2016; Fedele et al.
2017), relating these to an accurate surface density pro-
file requires several assumptions about the local physical
and chemical conditions to be made.
Recently, Teague et al. (2018a) demonstrated a tech-
nique to measure highly precise rotation velocities in
Corresponding author: Richard Teague
[email protected]
axisymmetric disks. For a geometrically thick disk with
gradients in temperature and density the rotation veloc-
ity is given by,
v2
rot
r
=
GM(cid:63)r
(r2 + z2)3/2
+
1
ρgas
∂P
∂r
,
(1)
where P = ngaskT is the gas pressure and ρgas is the
gas density. Thus, in combination with measurements
of the emission surface (Pinte et al. 2018a), deviations
from Keplerian rotation can be used to infer the pres-
sure gradient. Teague et al. (2018a) used this to place
tight constraints on the gas surface density profile of
HD 163296 and infer the presence of two Jupiter-mass
planets.
Others have also advocated the use of kinematics to
identify potential sources for changes in the pressure gra-
dient. Pinte et al. (2018a) reported kinematic evidence
of a wide separation ∼ 2MJup planet at ≈ 260 au in
HD 163296, extending far beyond the continuum edge.
Similarly, P´erez et al. (2018) showed that planet-disk in-
teractions will drive large non-Keplerian velocities which
can be used to locate potential perturbers.
In addition to searching for the signs of embedded
protoplanets, constraints of the pressure gradient are
2
Teague et al.
invaluable for interpreting observations of the dust. The
inward radial motion of particles due to the headwind
from gas rotating at sub-Keplerian speeds, as the gas
is supported by the radial pressure gradient, can very
rapidly deplete the disk of dust. To slow this depletion,
pressure bumps are frequently invoked, resulting in the
trapping of particles and thus extending the lifetime of
the dust disk (Pinilla et al. 2012). However, despite
the necessity of such pressure traps, direct evidence of
changes in gas pressure (rather than local enhancements
of dust interpreted as a dust trap) are lacking.
In this paper we present an improved method to mea-
sure the rotational velocity of a protoplanetary disk
which relaxes assumptions of the azimuthal symmetry
and intrinsic Gaussian line profiles which is described in
Section 2. In Section 3, we apply this method to archival
ALMA data of AS 209 and present a discussion of the
observed features. We summarise the findings and con-
clude in Section 5.
2. MEASURING THE ROTATION VELOCITY
In Teague et al. (2018a) we presented a method to
measure the rotation velocity of an axisymmetric disk.
The method required the minimization of the width of
the averaged line profile at a given radius after account-
ing for the projected rotation. While this method proved
to be robust, it makes the assumption that the result-
ing averaged profile would be a single Gaussian compo-
nent. We have improved upon this technique to relax
this assumption by modelling the stacked spectrum as
a Gaussian Process which allows a much more flexible
model (Foreman-Mackey et al. 2017). In this section we
review the method and describe the updates.
We make use of the fact that protoplanetary disks
are predominantly azimuthally symmetric. Line emis-
sion arising from the same radial location in the disk
should therefore be tracing the same physical and chem-
ical properties and thus possess the same profile. The
only difference will be in the line center which will be
offset from the systemic velocity by vrot · cos(θ) where θ
is the polar angle measured from the red-shifted major
axis. Note that this is not the polar angle measured in
the sky plane but must be calculated taking into account
the disk geometry (see the radial dotted lines in Fig. 1).
Figure 1 shows a toy model as an example. In panel
(a) the projected line-of-sight velocity is shown by the
filled contours. As the disk is flared, we see both the
top side and far side of the disk and this demonstrates
the need to correctly account for the emission height
of the disk. The dotted lines trace lines of constant r
and θ. Spectra extracted at the black dots, as shown
in the panels (b), (c) and (d), will be the same shape,
but offset on the velocity axis (shown in black). If vrot is
known, these spectra can be shifted back to the systemic
velocity, ready to be stacked, as shown by the gray lines.
This technique has been used previously to significantly
boost the SNR of spectra and increase sensitivity in the
outer edge of the disk (Teague et al. 2016; Yen et al.
2016; Matr`a et al. 2017) and a similar method used by
Yen et al. (2018) to measure dynamic stellar masses.
Rather than assuming vrot a priori we can use the
deprojected spectra to infer what the correct value is.
In Teague et al. (2018a) we used the linewidth of the
stacked spectrum as a proxy.
If the lines are depro-
jected using an incorrect value, the line centers will still
have some offset leading to a broadening of the final
line profile. Thus, the vrot value which minimizes the
final linewidth is the correct value as this is when all the
spectra are correctly aligned. However, this assumes
that the final profile is Gaussian which is often not the
case. For example, optically thick lines, such as 12CO
and potentially 13CO, will have line profiles which devi-
ate significantly from a Gaussian due to the saturation
of the line core.
Here we argue that a more flexible approach is to
model the stacked spectrum as a Gaussian Process (es-
sentially a probabilistic approach to modelling smooth,
non-parametric functions) and find the value of vrot
which minimizes the variance in the residuals. A simi-
lar approach has been used by Czekala et al. (2017) to
model the spectra of binaries.
An example of this approach is shown in Figure 2
where the left column shows the situation where an in-
correct vrot value has been assumed for the deprojec-
tion and thus there is significant scatter in the resid-
ual around the line wings, as shown in panel (b). Con-
versely, when the correct vrot has been used, as shown
in the right column, the scatter in residuals is near con-
stant across the profile. This allows for the case of highly
non-Gaussian profiles or if there is significant differences
in line brightness as a function of azimuth.
Figure 2 also demonstrates why a precision well below
the velocity resolution can be achieved. As the line pro-
file is sampled at different locations due to shifts from
the rotation, the deprojected spectra result in a sam-
pling of the intrinsic line profile at a factor of ∼ 10
higher. The gray bars in the residual plots show the
width of a single channel and demonstrate that the pro-
file is sampled at a much higher frequency after the shift.
Although some broadening may be present due to the
Hanning smoothing applied in the correlator, the result-
ing vrot will be insensitive to this as the optimization
does not care about the properties of the stacked line
profile other than it is smooth.
3
Figure 1. Demonstration of the method with a toy model. Panel (a) shows an example for the projected velocity arising from
a flared disk structure. The dotted lines show lines of constant radius and polar angle. Panels (b), (c) and (d) show spectra
extracted at the marked locations in black. The gray lines show the same spectra but shifted by −vrot · cos(θ), back to the
systemic velocity of the disk and thus able to be stacked.
Figure 2. Demonstration of the quality of fit method used. Each point represented a channel in a spectrum given some velocity
shift. The blue lines are the Gaussian Process (GP) model of these data and the gray solid line shows the true intrinsic line
profile. The bottom row shows the residual between the individual points and the GP model. If an incorrect vrot is used to
deproject the data (left column) significant variance is seen in the residuals. Conversely, when the correct vrot is used (right
column), the variances in the residuals is minimised. The shading in the residual plots shows the channel width, demonstrating
that after the deprojection of all the spectra, we sample the intrinsic line profile at a much higher rate than the observations,
allowing us to achieve a much higher precision on vrot than the native channel width.
−2−1012Offset(arcsec)−2−1012Offset(arcsec)(a)θ=0(b)v0=vLSR+vrot·cos(θ)θ=π/2(c)−3−2−10123Velocity(kms−1)θ=±π(d)−4−2024LineOfSightVelocity(kms−1)0.00.51.0Intensity(a)IncorrectRotationVelocityDeprojectedSpectraGPModelTruth0.00.51.0Intensity(c)CorrectRotationVelocityDeprojectedSpectraGPModelTruth−1000−50005001000Velocity(ms−1)−101Residual(b)−1000−50005001000Velocity(ms−1)−101Residual(d)4
Teague et al.
In practice this is performed using the Python package
celerite (Foreman-Mackey et al. 2017). This approach
has the added advantage that a more robust uncertainty
can be derived for vrot. As celerite naturally consid-
ers correlations between pixels, any spatial correlations
arising from the imaging (such as those described in Fla-
herty et al. 2018) will be accounted for in the uncer-
tainty. Furthermore, this approach more readily allows
for the inclusion of priors for vrot, or a more specific
model for the noise (for example the correlated noise
described in Teague et al. 2018b).
We have tested this method with various levels of
noise, velocity resolutions, azimuthal asymmetries in
both line width and peak and non-Gaussian lines profile.
A thorough examination of the accuracy and precision
achieved is presented in Appendix A, including the im-
pact of azimuthal structure in the line profiles or when
the near and far sides of the disk are spatially resolved.
In brief, however, the accuracy achieved by this method
when there is no strong azimuthal structure can be es-
timated by
(cid:19)−1.12±0.01
(2)
(cid:18)(cid:114) 2πr
44.4 ± 0.1 m s−1 ×
· SNR
10
θbeam
where SNR is the signal-to-noise achieved in a single
pixel at radius r and θbeam is the beam FWHM. Az-
imuthal deviations in the linewidth and peak do not sig-
nificantly hinder this approach unless they are greater
than ≈ 50% in magnitude.
Examples of the code used for this and Jupyter Note-
books containing guides on how to use them can be
found at https://github.com/richteague/eddy. Version
1.0 of the code used for this paper can be obtained from
https://doi.org/10.5281/zenodo.1440051.
3. THE ROTATION CURVE OF AS 209
To demonstrate this method we use archival data of
AS 209 (2015.1.00486.S, PI Fedele, D.). Continuum ob-
servations from this project have been previously pre-
sented in Fedele et al. (2018) which have suggested that
multiple gaps in the continuum can be driven by a single
planet. Here we focus only on 12CO emission, leaving
a thorough analysis of the three CO isotopologues and
DCO+ to be presented in a future work (Favre et al., in
prep.).
3.1. Observations
Data reduction followed the same process as outlined
in Fedele et al. (2018). The data were calibrated using
the provided scripts in casa v4.4 before moving to casa
v5.2 for the imaging and self-calibration. Phase gain
tables were calculated on the continuum window then
applied to the three spectral line windows containing
12CO emission.
We consider two cases, both with and without the con-
tinuum subtracted from the line data. The continuum
is removed using the task uvcontsub which linearly in-
terpolates the the continuum from line-free channels in
the uv-plane. As discussed in Boehler et al. (2017), this
can lead to an under-estimation of the true total inten-
sity of the line as the molecular gas will absorb some of
the continuum. While this will affect the inferred tem-
perature or column density which require absolute flux
measures, our method is insensitive to such effects. As
the effect is essentially independent of frequency, at least
across the line profile, see for example Fig. 8 in Boehler
et al. (2017), then the change in the line profile will be
symmetric about the line center and thus not change
our derivation of velocity.
Imaging the continuum emission we used uniform
weighting resulting in a beamsize of 0.15(cid:48)(cid:48)×0.13(cid:48)(cid:48) at a po-
sition angle of 2.6◦. A RMS noise of σ = 65 µJy beam−1
was measured in continuum free regions of the contin-
uum map and an integrated intensity of 251 mJy was
measured, consistent with previous observations ( Oberg
et al. 2011; Huang et al. 2016).
We perform a different approach for the 12CO emis-
sion. Using the tclean task, we first image the emis-
sion with natural weighting, to maximize sensitivity, and
with a square 3(cid:48)(cid:48) × 3(cid:48)(cid:48) box as the mask. From this image
we generate a first moment map clipping values below
2σ where σ was measured in a line free channel. To this
first moment map we fit a Keplerian profile to derive
a position angle, vLSR and Mstar, holding the inclina-
tion constant at i = 35.3◦ (Fedele et al. 2018) in order
to break the M · sin i degeneracy. These values were
then used to generate a Keplerian mask (masking out
regions where, given the derived rotation pattern, we
would not expect emission to arise) for the data which
was convolved with the beam and checked to encompass
the whole emission. The data were then imaged and
CLEANed again using this Keplerian mask.
For precise measurements of vrot, sensitivity is more
important than spatial resolution. Therefore we use nat-
ural weighting for the imaging yielding beam sizes of
0.23(cid:48)(cid:48) × 0.19(cid:48)(cid:48) at a position angle of −79.3◦. The data
were imaged at a velocity resolution of ≈ 160 m s−1
(≈ 244 kHz). The RMS noise in a line free channel was
3.3 mJy beam−1 and we measure an integrated flux of
7.99 Jy km s−1, consistent with previous measurements
(Huang et al. 2016).
Figure 3 summarises the observations, showing the ve-
locity integrated fluxes (using a 2σ clip and the CLEAN
5
Figure 3. Observations of 12CO and the 233 GHz continuum. The left two panels show the integrated intensities while the
right panel shows the normalised azimuthally averaged profile. Synthesized beamsizes are shown in the bottom left of each
panel. The integrated flux is shown in the top left. The coloured contours have been saturated in the centre to highlight the
extended emission. The radial profiles are sampled at quarter beam spacing and the error bars show the standard deviation of
the annulus. 12CO emission is split into 'unobscured' and 'obscured', east and west of the dotted lines (at ±55◦ either side of
the major axis) shown in the left most panel.
mask), and their radial profiles. Significant absorption
is seen in the west half of the 12CO emission, likely due
to cloud contamination at vLSR (cid:46) 5 km s−1 ( Oberg
et al. 2011; Huang et al. 2016). Assuming that for
vLSR (cid:38) 5 km s−1 the emission is free from absorption,
ratios of the intensity profiles suggest that the cloud ab-
sorbs ≈ 30% of the 12CO emission on the Western side
of the disk.
There are no clear deviations from a Keplerian pattern
observed in the channel maps indicative of large scale
kinematic features (Perez et al. 2015; Pinte et al. 2018b),
however the cloud contamination and limited sensitivity
may limit the visibility of such features.
3.2. Measuring Rotation Curves
In order to use the method presented in Section 2 an
initial estimate of the rotation profile and the emission
height, in order to properly deproject the data into an-
nuli of constant radii are required.
For an initial estimate of the expected rotation pro-
file we fit a Keplerian pattern, including a conical emis-
sion surface (Rosenfeld et al. 2013) using the MCMC
ensemble sampler emcee (Foreman-Mackey et al. 2013).
We calculate a map of the line-of-sight velocities using
the method presented in Teague & Foreman-Mackey
(2018) which fits a parabola to the pixel of peak intensity
and its two neighbouring pixels. This method allow us
to discriminate between emission arising from the near
side of the disk and the far side, while achieving a sub-
channel precision measurement of the line centroid. In
addition, the cloud absorption will less strongly bias the
measurement of the maximum coordinate. The emis-
sion surface is calculated as z = r · tan ψ where ψ is
the angle between the disk midplane and the emission
surface1.
In addition, with each call of the likelihood
function the rotation pattern was convolved with a 2D
Gaussian matching the synthesized beam for each ob-
servation to account for convolution effects in the inner
regions of the disk (Walsh et al. 2017).
The best-fit rotation pattern, assuming a fixed in-
clination i = 35.3◦, was described with M(cid:63) = 1.16 ±
0.01 Msun, PA = 86.7◦ ± 0.1◦, vLSR = 4670 ± 5 m s−1
and ψ = 13.1◦ ± 0.3◦. The uncertainties are the stan-
dard deviation of the posterior distribution. The in-
ferred stellar mass is slightly larger than the previously
found Mstar = 0.9 Msun (Andrews et al. 2009), likely
due to better resolving the emission in high velocity
channels which contain the most information to distin-
guish between stellar masses. The position angle and
systemic velocity are consistent with previous determi-
nations (Andrews et al. 2009; Huang et al. 2016).
Using the method presented in Pinte et al. (2018a)
to measure the emission height we find good agreement
with a conical model with ψ ≈ 13◦, consistent with the
determination from the ninth moment map fitting. As
the emission surface determination relies on the asym-
metry across the major axis of the disk, cloud absorp-
1 We have also tried a more complex surface of z = z0 × rφ,
however the spatial resolution of the data meant that the fits were
unable to converge.
−2−1012Offset(arcsec)−2−1012Offset(arcsec)7.99Jy12CO(2−1)−2−1012Offset(arcsec)251mJy233GHz0100200300Radius(au)0.00.20.40.60.81.0NormalisedIntensityUnobscured12COObscured12CO233GHz050100150200mJybeam−1kms−10369mJybeam−10.00.51.01.52.02.5Radius(arcsec)6
Teague et al.
tion will not impact this result as this only results in
an asymmetry across the minor side of the disk. Off-
sets in the ellipse centres were found in the northern
direction meaning that the southern side of the disk is
closer consistent with the preferred orientation proposed
by Avenhaus et al. (2018) who showed that features in
the NIR scattered light better align with continuum fea-
tures when the southern side, rather than the northern
side is closer. This orientation is in addition consistent
with that found from the fitting of the ninth moment
map.
Following the method in Teague et al. (2018a) and
with the modification described in Section 2 we measure
an azimuthally averaged rotation profile. We consider
three cases: firstly we deproject the sky-plane coordi-
nates into disk coordinates to correctly account for the
flaring, secondly we consider a geometrically thin disk,
and finally we consider the flared disk with no contin-
uum subtraction. For each we we then bin the data into
annuli with a width of a quarter of the beam (≈ 0.05(cid:48)(cid:48))
and derive a vrot value. We further model the radial
profile as a Gaussian Process, requiring the profile to
be smoothly varying and to account for the correlations
due to the sub-beam size sampling.
These rotation profiles are shown in Figure 4a, with 1σ
uncertainties. The dotted line is a fit of a geometrically
thin Keplerian rotation profile assuming i = 35.7◦ and
resulting in Mstar = 1.25 Msun. This profile is used as
the reference profile for the derived values, however we
stress that this should not be taken at the true stellar
mass as we are unable to disentangle large scale effects of
the pressure gradient from the stellar mass. The middle
and bottom panels show the residual in m s−1 and as a
percentage, respectively. Gray and black lines show the
locations of rings observed in the mm continuum and
NIR scattered light, respectively.
All three scenarios yield comparable radial vrot profiles
which are consistent within 3σ of one another. This sug-
gests that continuum subtraction does not significantly
affect the derived rotation profile.
When plotting residuals from a Keplerian profile, as in
panels (b) and (c), large negative gradients are indica-
tive of a pressure maximum, while large positive gradi-
ents are indicative of pressure minima (for an example,
see Fig. 1 of Teague et al. 2018a). We see that the two
inner rings for both mm and near-infrared (NIR) emis-
sion are centred on pressure maxima, as predicted from
grain evolution models (Birnstiel et al. 2012; Pinilla et
al. 2012). The outer ring in scattered light, however,
appears at the outer edge of a large pressure minimum
centred at ≈ 250 au. This will be discussed in the fol-
lowing section.
4. A PERTURBED PHYSICAL STRUCTURE
As the deviations from a smooth rotation curve can
be driven by changes in local temperature, density and
the height of the emission, it is hard to isolate the main
driver.
In this section we use a toy model which we
perturb in order to reproduce the observed deviations
in rotation velocities and infer the underlying pressure
profile.
4.1. The Toy Model
The model is based on the commonly used prescription
using a simple physical structure (Rosenfeld et al. 2013;
Williams & Best 2014), with specific values taken from
Huang et al. (2016). The total gas surface density is
given by Lynden-Bell, & Pringle (1974),
(cid:18) r
(cid:19)−γ
r0
(cid:32)
(cid:21)2−γ(cid:33)
(cid:20) r
r0
exp
−
Σgas(r) = Σ0
(3)
where r0 = 100 au, γ = 1 and Σ0 = 4.95 g cm−2 such
that the total disk mass is 0.035 Msun, a factor of 100
times larger than the dust mass used in Fedele et al.
(2018). This is inflated to a volume density assuming a
Gaussian density profile,
(cid:18)
(cid:19)
− z2
2Hp(r)2
√
Σgas(r)
2πHp(r)
ρgas(r, z) =
(4)
with the scale height parametrized as Hp = 10 ×
(r/r0)1.26 au.
exp
The thermal structure follows the prescription in Dar-
tois et al. (2003) which smoothly connects two bound-
ary layers, the midplane and the atmosphere, through a
trigonometric function,
Tatm
Tatm +(cid:0)Tmid − Tatm
T =
(cid:1) cos2δ(cid:16) zπ
2zq
(cid:17)
z ≥ zq
z < zq
(5)
where δ = 2 and zq = 4Hp. The midplane and atmo-
spheric temperatures are also described by radial power-
laws, Tmid(r) = 15.7 × (r/r0)−0.48 K and Tatm(r) =
47.4 × (r/r0)−0.50 K, respectively.
Following Huang et al. (2016), we consider a homo-
geneous distribution of CO throughout the disk with-
out taking into account the freeze-out or photodissocia-
tion. As the model in Huang et al. (2016) only considers
CO rather than H2, we chose a relative abundance of
x(CO) = 1.7 × 10−6 in order to recover their prescribed
column density profiles.
7
Figure 4. Residuals between the measured vrot and the Keplerian rotation curve found from fitting the first moment map. Error
bars show 1σ uncertainties with blue assuming a flared surface for the deprojection, dark gray bars assuming a geometrically
thin disk and light gray bars a flared disk without continuum subtraction. The vertical lines show the centre of the rings in
dust mm emission (gray, Fedele et al. 2018) and scattered light (black, assuming that the southern side is closer to the observer,
Avenhaus et al. 2018). The synthesised beamsize is shown in the top left corner.
4.2. Perturbations
We only consider the emission region of the 12CO
which should be narrow in the vertical direction due
to the high optical depth of the line. From the model
we find that the 12CO contribution function weighted
height, temperature and gas density are well described
by power laws over the region of interest, 30 au ≤ r ≤
320 au:
r
100 au
T12CO(r) = 41 K ×(cid:16)
n12CO(r) = 9.6 · 106 cm−3 ×(cid:16)
z12CO(r) = 23 au ×(cid:16)
,
(cid:17)−0.57
(cid:17)−2.29
(cid:17)1.04
.
r
100 au
r
100 au
Taking each of these quantities in turn, we model a
perturbation vector as a sum of six Gaussian curves
and multiply the power-law describing that property by
this to create a perturbed profile (similar to the pertur-
bations used to model the continuum intensity profile
Fedele et al. 2018). Using this perturbed profile and
fixing the other two properties, vrot is calculated using
Eqn. 1 and compared to the observed vrot profile. Al-
though in reality these three properties are highly cou-
pled, this approach allows us to quantify the extreme
cases which are consistent with the data.
The resulting best-fit perturbed profiles are shown in
Fig. 5. For comparison, the brightness temperature, a
proxy of the local temperature for optically thick lines,
of the 12CO emission from the cloud-free region is shown
in panel (b) and the derived 12CO emission surface is
shown in panel (c), both with blue error bars. The error
bars on the perturbed model represents the scatter of
200 draws from the MCMC fitting.
(6)
,
(7)
(8)
The subscript 12CO is to show that these profiles trace
the 12CO emission region and do not necessarily trace a
fixed height in the disk.
1.01.52.02.53.0vrot(kms−1)(a)FlaredFlatWithContinuum−100−50050100vrot−vkep(ms−1)mmmmNIRNIRNIRNIRbeam(b)50100150200250300Radius(au)−505δvrot(%)mmmmNIRNIRNIRNIRbeam(c)0.51.01.52.02.5Radius(arcsec)8
Teague et al.
Changes in the emission height, shown in panel (c),
require larger deviations, particularly between 150 and
250 au, than are allowed from observations. In addition,
these perturbations would place the continuum rings
centred at local minima in height, shielding the rings
from direct irradiation from the star, as shown by the
gray shaded region. Such deviations could not be pos-
sible given the structure observed in the scattered light
(Avenhaus et al. 2018).
We therefore conclude that the deviations in the ro-
tation velocity are likely driven by change in the radial
gas pressure gradient, a combination of both density and
temperature.
4.3. Pressure Profile
It has been a long standing assumption that grains
collect in pressure maxima resulting in environments
conducive to grain growth and the beginnings of planet
formation (Whipple 1972; Birnstiel et al. 2012; Pinilla
et al. 2012). This is because the velocity of the grains
relative to the gas is given by udrift ∝ ∂P / ∂r (Weiden-
schilling 1977), then a particle will drift towards pressure
maximum. Thus, as grains are predominantly confined
within traps this significantly slows radial drift. With
constraints on the pressure gradient, we are able to di-
rectly test this assumption.
From the perturbed radial profiles of T12CO and n12CO
we can calculate the pressure profile traced by the 12CO
emission. Figure 6 compares this inferred pressure pro-
file and the derivative of its logarithm with the radial
continuum emission profile and the r2-scaled scattered
light intensity from Avenhaus et al. (2018), assuming
the southern side of the disk is closest. We see a slight
offset in the radial location of the rings in mm and µm
sized grains, with a better match to the pressure maxima
with the µm sized particles.
The absolute scaling of this pressure profile is depen-
dent on the density assumed for the disk model (Eqn. 7).
Changes in the density structure will result in different
amplitude perturbations. Despite this degeneracy, the
location of the perturbations will remain constant.
Due to the limited resolution of these observations
(≈ 29 au for the line emission) the velocity features are
not resolved and so will underestimate the true depth.
Future, higher resolution observations of the line emis-
sion will better constrain the depth of these pressure
perturbations and thus their gradient. Such observa-
tions will be essential in constraining the level of particle
trapping in such pressure maxima.
Only the outer most ring at ≈ 250 au does not coincide
with a pressure maxima, rather with a pressure minima.
One possible interpretation of this is that such a drop in
Figure 5. Perturbed physical profiles resulting in the ob-
served vrot. The red dotted lines show the fiducial models
while the red error bars show the 16th to 84th percentile
range from 200 draws from the posterior distributions. The
blue error bars show the observations: 12CO peak brightness
temperature for the temperature and the inferred emission
surface. Note there are no observable constraints on density.
Gray and black vertical lines show the location of the rings
in mm continuum and NIR scattered light, respectively. In
panel (c) the shaded region shows the shadowed region for
the perturbed model showing that many of the NIR scattered
rings would be located in shadow.
Decreases in the density are found coincident with the
gaps observed in the mm continuum centred at 62 and
103 au (Fedele et al. 2018). Additionally, a peak is ob-
served at ≈ 150 au, consistent with the required excess
used to explain the ringed CO isotopologue emission
(Huang et al. 2016).
Large changes in temperature as shown in panel (b)
can be ruled out by the line emission as TB ≤ Tgas
(note however that these temperatures may be under-
estimated, particularly over the dust rings, due to over
subtraction of the dust continuum). Analysis of alterna-
tive transitions of 12CO will help constrain the temper-
ature structure within the inner disk and provide limits
for possible changes in temperature, while higher angu-
lar resolution will allow for small, local changes to be
resolved.
105106107108Density(cm−3)(a)mmmmNIRNIRNIRNIRFiducialPerturbedObservations0306090Temperature(K)(b)mmmmNIRNIRNIRNIR050100150200250300Radius(au)0306090Height(au)ShadowedRegion(c)mmmmNIRNIRNIRNIR0.00.51.01.52.02.5Radius(arcsec)9
grains in the disk atmosphere, there is a vertical de-
pendence in the location of this maxima. This is consis-
tent with the radial offsets found in the location of pres-
sure minima traced at different heights in HD 163296
(Teague et al. 2018a) and with similar features seen in
three-dimensional simulations (see, for example, Fig. 4
of Fung, & Chiang 2016).
However, it is unclear whether small particles can as
efficiently trapped at higher altitudes than larger, mm-
sized particles in the midplane. As the particles become
trapped, the rate of collision increases and grain growth
is hastened. Larger grains will rapidly settle towards
the midplane and drift radially inwards (Dullemond, &
Dominik 2004). Two-dimensional simulations combined
with accurate vertical profiles for the pressure gradient
will be required to properly test this claim.
5.2. Sources of Perturbations
To account for the observed deviations in velocity we
require at least 3 significant perturbations to the physi-
cal structure of the disk at approximately 50 au, 100 au
and 250 au. The most attractive scenario for the source
of these perturbations is the presence of planets. Fedele
et al. (2018) demonstrated that the continuum emis-
sion profile can be explained either by a Saturn mass
planet at 95 au and potentially a second planet less than
0.1 MJup at 57 au. The density contrast required to re-
cover the rotation velocities are broadly consistent with
thosed used to model the continuum emission profile.
The pressure minima at ∼ 250 au is less likely to be
opened by a planet. Dynamical time scales at these
radii are prohibitive for core formation and the forma-
tion of planets, although planet formation via gravita-
tional instability is a possibility (Boss 1997). Recently,
Pinte et al. (2018b) found similar kinematic signatures
for a ∼ 2 MJup planet at 260 au in the disk around
HD 163296. This suggests that there may be a popula-
tion of massive, wide separation planets which contin-
uum observations are not sensitive to.
Aside from a planetary origin, other hydrodynamic
instabilities have been shown to result in similar pertur-
bations to the disk physical structure. The magneto-
rotational instability (MRI) has been shown to drive
large gaps in the gas surface density at the outer edge
of the dead zone (Flock et al. 2015). However, esti-
mates of the dead-zone extend only reach out to ≈ 60 au
(Cleeves et al. 2015), far further in than the observed de-
viation. Higher resolution observations of molecular line
emission will help constrain the local physical properties
where perturbations are observed and help distinguish
between possible sources.
Figure 6. Top: Radial profiles of the mm continuum (blue)
and the µm scattered light (red, from Avenhaus et al. 2018,
and scaled by r2 to account for the drop in incident pho-
tons). The vertical dashed lines show the location of peaks
in these profiles. Middle: samples of the inferred pressure
profiles consistent with the measured vrot values. Bottom:
The pressure gradient showing the peaks of the scattered
light profile align with the pressure maxima.
pressure will result in decrease of the disk scale height.
If this is only a shallow perturbation, such that the far
side is not shadowed, then the outer wall of this dip
will have a larger angle of incidence for stellar light and
thus scatter more effectively from the sub-µm grains,
resulting in a ring despite the lack of a pressure maxima.
5. DISCUSSION
We have demonstrated that we are able to use gas
kinematics to infer the presence of perturbations in the
physical structure of AS 209 and infer the radial pres-
sure profile. As these constraints are free of assumptions
about the line excitation, they are hugely complimen-
tary to traditional methods aiming to recover emission
morphology.
5.1. Vertical Dependence of Pressure Traps
Better correlation is found between the pressure max-
ima and the rings observed in NIR scattered light than
the the mm continuum. This suggests that if the rings
are due to pressure confinement of the sub-µm sized
0.00.51.0NormalisedIntensityScatteredLightContinuum−22−20−18lnP50100150200250300Radius(au)−6−303dlnP/dlnr0.51.01.52.02.5Radius(arcsec)10
Teague et al.
5.3. CO Desorption Front
Huang et al. (2016) interpreted the ringed structure
observed in C18O emission (and tentatively in 13CO)
at 150 au as an enhancement in the local abundance
of CO due to a desorption front. The authors argued
that the lower opacity of the disk in regions beyond the
mm continuum edge would allow for more efficient non-
thermal desorption processes to occur resulting in a lo-
cal enhancement in CO abundance as hypothesized by
Cleeves (2016).
As shown in Section 4, we require enhancements in
the H2 gas pressure at 150 au in order to explain the ve-
locity structure, either through an increase in density or
temperature. Such changes can also explain the increase
in the optically thin line emission without the need for
an enhanced in the local CO abundance. However, such
changes in the physical structure are likely to also affect
the chemistry with an increase in temperature leading
to more thermal desorption and higher CO column den-
sities.
6. CONCLUSIONS
We have extended the method presented in Teague et
al. (2018a) for measuring rotational velocities to allow
for non-Gaussian line profiles and for objects with signif-
icant azimuthal structure. Application of this method
to archival data of AS 209 revealed persistent deviations
from a smooth Keplerian profile.
Using a toy model of AS 209, we are able to quan-
tify the deviations required in the temperature, density
and height of the emission to match the observed per-
turbations resulting in deviations of up to 80%. Future
work using models with self-consistent physical struc-
tures will be able to disentangle the relative contribu-
tions from the density and temperature terms. Compar-
ison of the resulting pressure profiles provides evidence
for the pressure trapping of sub µm particles in the disk
atmosphere, while a radial offset in the ring locations
for the mm continuum and the scattered NIR light sug-
gest that the location of the pressure minimum moves
radially outwards at higher altitudes in the disk.
A perturbation in the disk structure is inferred at
≈ 250 au, far beyond the edge of the mm-continuum,
resulting in deviations of up to 5% from Keplerian ro-
tation. A planetary origin for this object is unlikely as
the dynamical time scales at such large radii make the
initial stages of core accretion inefficient.
This work demonstrates the utility of studies of the
gas kinematics and the ability to provide unique con-
straints for the interpretation of high angular resolution
continuum observations.
We thank the referee for their helpful comments which
have made the presentation of this method much clearer.
This paper makes use of the following ALMA data:
JAO.ALMA#2015.1.00486.S. ALMA is a partnership of
European Southern Observatory (ESO) (representing its
member states), National Science Foundation (USA),
and National Institutes of Natural Sciences (Japan), to-
gether with National Research Council (Canada), Na-
tional Science Council and Academia Sinica Institute of
Astronomy and Astrophysics (Taiwan), and Korea As-
tronomy and Space Science Institute (Korea), in coop-
eration with Chile. The Joint ALMA Observatory is
operated by ESO, Associated Universities, Inc/National
Radio Astronomy Observatory (NRAO), and National
Astronomical Observatory of Japan. The National Ra-
dio Astronomy Observatory is a facility of the National
Science Foundation operated under cooperative agree-
ment by Associated Universities, Inc. This project has
received funding from the European Research Council
(ERC) under the European Unions Horizon 2020 re-
search and innovation programme under grant agree-
ment No 714769 and was supported by funding from
NSF grants AST-1514670 and NASA NNX16AB48G.
R.T. would like to thank Bertram Bitsch, Kees Dulle-
mond and Mario Flock for insightful discussions and
Henning Avenhaus for sharing the scattered light data.
Software:
bettermoments (Teague & Foreman-
Mackey 2018), CASA (v4.2 & v5.2, McMullin et al. 2007),
celerite (Foreman-Mackey et al. 2017), eddy (Teague &
Birnstiel 2018),emcee(Foreman-Mackeyetal.2013),mat-
plotlib (Hunter 2007), numpy (van der Walt et al. 2011),
scipy (Jones et al. 2001)
APPENDIX
A. RECOVERING THE ROTATION VELOCITY
In this Appendix we demonstrate the robustness of the derived vrot and quantify the precision which can be achieved
with this method. The code used to calcaulted vrot and the model spectra can be found at https://github.com/
richteague/eddy. Version 1.0 of the code used for this paper can be obtained from https://doi.org/10.5281/zenodo.
1440051.
11
Figure 7. Accuracy and preicision achieved for the method described in Section 2 while varying the signal-to-noise of the
spectra, the linewidth of the line and the number of spectra used. The blue points each represent a random sample, while the
solid like shows the median of the distribution and the dotted show the 16th and 84th percentiles (equivalent to one standard
deviation for a Gaussian distribution).
A.1. General Properties
We first consider the case of well behaved data: intrinsic Gaussian profiles with Gaussian noise. To model the 12CO
emission we generated 20, 000 sets of model data. Each sample contained a random number of spectra, N ∈ [6, 60],
linearly spaced across the 2π azimuth. This range encompasses the expected number of independent beams for disk
observed with ALMA at ∼ 0.1(cid:48)(cid:48) resolution: N ≈ 2πr / θbeam.
The underlying profile was assumed to be Gaussian described by TB ∈ [5, 40] K, ∆V ∈ [100, 400] m s−1 and
vrot ∈ [0.5, 3.5] km s−1. Each was then corrupted by Gaussian noise to achieve a SNR ∈ [2, 20]. These values were
chosen to represent typical line properties observed in protoplanetary disks. They were calculated on a velocity axis
with a resolution of 160 m s−1 meaning that the FWHM of the line was sampled between roughly 1 and 4 times.
For each sample of lines, we inferred vrot following the method described in Section 2. The difference from the true
value are shown in Fig. 7. As expected, the accuracy achieved increases with the SNR of the data and the number of
lines used as both of these increase the SNR of the stacked spectra. Marginalizing over all intrinsic line properties we
can model the accuracy of this method via the power-law it to the 16th to 84th percentile range of residuals (shown
by the dotted lines in Fig. 7) as,
Accuracy = 44.4 ± 0.1 m s−1 ×
,
(A1)
showing that with SNRs readily achievable by ALMA (due to both the overall sensitivity and the small beam sizes
achieved in order to better sample the annulus), accuracies of a few meters per second are possible. A measure of the
precision of the results can be estimated by the width of the posterior distribution for vrot for each case. This is well
fit with the profile,
(cid:18)(cid:114) 2πr
· SNR
10
θbeam
(cid:19)−1.12±0.01
(cid:18)(cid:114) 2πr
· SNR
10
θbeam
(cid:19)−1.39±0.04
Precision = 23.9 ± 0.6 m s−1 ×
,
(A2)
which is roughly a factor of two larger than the accuracy. Thus the assumed 3σ uncertainties quoted in this work (and
Teague et al. 2018a), are consistent with the true vrot value.
A.2. Azimuthal Asymmetry
For the case of AS 209 there is strong azimuthal asymmetry due to the cloud absorption.
In this section we
demonstrate how such azimuthal structure in either TB or ∆V affects the inferred vrot. Similar to the previous examples
we generate model spectra, however reduce the parameter space by considering only ∆V = 300 m s−1, N = 20 and
SNR = {5, 10, 15}. We then include a periodic perturbation in ∆V , TB or both parameters, parameterised as
(cid:18) θ + χ
(cid:19)
f
δ = 1 + δ0 · sin
(A3)
5101520Signal-to-NoiseRatio−100−50050100Residual(ms−1)(a)1.52.02.53.03.54.0SamplesperFWHM(b)102030405060NumberofIndependentBeams(c)12
Teague et al.
Figure 8. Accuracy of the Gaussian Process method when azimuthal structure is considered. Each call randomly select
{δ0, χ, f}, as described in Eqn. A3. The left panel shows when the deviation is only applied to the linewidth, the center panel
to the line peak and the right panel, both parameters.
where δ0 controls the strength of the deviation, χ ∈ [−π, π) is a random number to offset the deviation and f is an
integer frequency. For this Appendix we only consider f = {1, 2} and δ0 ∈ [0, 0.5]. For the case of AS 209, the cloud
contamination leads to a δ0 ≈ 0.3.
The results are shown in Fig. 8 where each panel shows 1, 100 samples. The dotted lines show the 16th and 84th
percentiles of the distribution and the solid line shows the median. These show that accuracy is not strongly affected
by the inclusion of azimuthal structure. This suggests that the method is able to robustly recover the an accurate
measure of the line to a an accuracy of (cid:46) 20 m s−1 even when both ∆V and TB have perturbations of up to 30%.
A.3. Spatially Resolved 3D Structure
With ALMA now able to routinely spatially resolve the near and far side of the disk for molecules with high emission
surfaces (de Gregorio-Monsalvo et al. 2013; Rosenfeld et al. 2013), emission from both the top and bottom sides of the
disk will be visible along a line of sight. This results in two components rather than a single component which could
potentially cause problems as demonstrated in Fig. 9.
To demonstrate the robustness of this technique against such contamination, we consider a simple model set-up.
Each line of sight will be the combination of a main Gaussian component from the near side of the disk and a slightly
offset secondary peak from the rear side of the disk. The secondary peak will have a smaller amplitude as this will be
tracing the snow-surface of the far side of the disk (Pinte et al. 2018a), thus TB ≈ 21 K.
To calculate the offset of the secondary peak we consider a disk with a given stellar mass, M(cid:63), inclination, i, and
emission surface described by z = z0 × rφ, where z0 = 0.3 and φ = 1.25. For a given radius we first calculate the
mapping of disk coordinates to sky coordinates via (see also Rosenfeld et al. 2013),
(cid:32)
(cid:33)
(cid:32)
xsky
ysky
=
(cid:0)ydisk − zdisk / sin(i)(cid:1) · cos(i)
xdisk
(cid:33)
.
(A4)
We then calculate where these sky coordinates intercept the far side of the disk. For the emission surface of the far
side of the disk we consider a smaller aspect ratio of z / r = 0.1 as this emission will arise from the snow surface,
as discussed previously, and thus be tracing a deeper (closer to the midplane) region. This results in two sets of
coordinates: (xdisk, ydisk, zdisk) for the front side of the disk and the same for the far side of the disk. The front side
coordinates will describe a ring of radius rdisk with height zdisk while the rear side will sample a range of rdisk values
and thus different zdisk and vrot values.
Figure 10a shows example spectra for disk matching the parameters of AS 209. The contamination from the rear
side of the disk is seen as small components offset from the main line centre with the largest deviations in the regions
between the major (θ = 0) and minor (θ = ±π) axes.
In Fig. 10b we show an annulus of spectra including the appropriate noise (σrms ≈ 2 K) and channel width (∆Vchan =
160 m s−1). The red lines show the line centers before deprojection. Figure 10b shows the deprojected spectra (in gray
dots) and the stacked profile as a solid line. As the rear side components are varying in their offset for each line, they
do not stack coherently and are thus lost in the noise, allowing for a good fit of vrot to be found even in this scenario,
achieving an accuracy of < 0.4%.
0.00.10.20.30.40.5δ0−100−50050100Residual(ms−1)(a)Perturbationsin∆V0.00.10.20.30.40.5δ0(b)PerturbationsinTB0.00.10.20.30.40.5δ0(c)Perturbationsin∆VandTB13
Figure 9. Demonstrating the origin of the double Gaussian line profiles for high spatial resolution data. The top panel shows
the spectrum extracted at (1.5(cid:48)(cid:48), 0.5(cid:48)(cid:48)), shown by the white dot in the channel maps. The velocities of the channel maps are
shown in the top right of each panel and as a vertical dotted line in the spectrum. For a given position we find two Gaussian
components due to the near and far sides of the disk with the near side being brighter.
In the case of very high signal-to-noise data with very small beam sizes, the far side components will become more
apparent. This can be circumvented using the initial method of minimizing the width of a Gaussian line profile as this
prior will be less sensitive to the contamination.
REFERENCES
Andrews, S. M., Wilner, D. J., Hughes, A. M., et al. 2009,
Czekala, I., Mandel, K. S., Andrews, S. M., et al. 2017,
ApJ, 700, 1502.
ApJ, 840, 49.
Avenhaus, H., Quanz, S. P., Garufi, A., etal. 2018, ApJ,
Dartois, E., Dutrey, A. & Guilloteau, S. 2003, A&A, 399,
863, 44.
773.
Birnstiel, T., Klahr, H. & Ercolano, B. 2012, A&A, 539,
Dullemond, C. P., & Dominik, C. 2004, A&A, 421, 1075.
A148.
Fedele, D., Carney, M., Hogerheijde, M. R., et al. 2017,
Boehler, Y., Weaver, E., Isella, A., et al. 2017, ApJ, 840, 60.
A&A, 600, A72
Boss, A. P. 1997, Science, 276, 1836.
Fedele, D., Tazzari, M., Booth, R., et al. 2018, A&A, 610,
Cleeves, L. I., Bergin, E. A., Qi, C., et al. 2015, ApJ, 799,
A24.
204.
Flaherty, K. M., Hughes, A. M., Teague, R., et al. 2018,
Cleeves, L. I. 2016, ApJ, 816, L21.
ApJ, 856, 117.
01234Velocity(kms−1)010203040BrightnessTemperature(K)(a)(b)(c)(d)(e)FarSideNearSide−202Offset(arcsec)0.9kms−1(b)NearSideFarSide1.6kms−1(c)024Offset(arcsec)−202Offset(arcsec)2.4kms−1(d)024Offset(arcsec)3.1kms−1(e)14
Teague et al.
Figure 10. Examples for a flared disk. The top panel shows example spectra including the contamination from the rear side
of the disk. The middle panel shows example spectra for an AS 209 like disk with parameters matching those described in
Section 3.1. The red lines indicate the line centres. The bottom panel shows the results of the deprojection showing that the
contamination from the far side does not significantly perturb the final stacked spectra.
Flock, M., Ruge, J. P., Dzyurkevich, N., et al. 2015, A&A,
McMullin, J. P., Waters, B., Schiebel, D., et al. 2007,
574, A68.
Astronomical Data Analysis Software and Systems XVI,
Foreman-Mackey, D., Hogg, D. W., Lang, D., et al. 2013,
Publications of the Astronomical Society of the Pacific,
125, 306.
Foreman-Mackey, D., Agol, E., Ambikasaran, S., et al.
2017, AJ, 154, 220.
Fung, J., & Chiang, E. 2016, ApJ, 832, 105.
de Gregorio-Monsalvo, I., M´enard, F., Dent, W., et al.
2013, A&A, 557, A133.
Huang, J., Oberg, K. I. & Andrews, S. M. 2016, ApJ, 823,
L18.
Hunter, J. D. 2007, Computing in Science and Engineering,
9, 90.
Isella, A., Guidi, G., Testi, L., et al. 2016, PhRvL, 117,
251101.
127.
Oberg, K. I., Qi, C., Fogel, J. K. J., et al. 2011, ApJ, 734,
98.
Perez, S., Dunhill, A., Casassus, S., et al. 2015, ApJ, 811,
L5.
P´erez, S., Casassus, S. & Ben´ıtez-Llambay, P. 2018,
MNRAS, 480, L12.
Pinilla, P., Birnstiel, T., Ricci, L., et al. 2012, A&A, 538,
A114.
Pinte, C., M´enard, F., Duchene, G., et al. 2018, A&A, 609,
A47.
Pinte, C., Price, D. J., M´enard, F., et al. 2018, ApJ, 860,
L13.
Jones, E., Oliphant, T., Peterson, P., et al. 2001, SciPy:
Open source scientific tools for Python.
Pohl, A., Benisty, M., Pinilla, P., et al. 2017, ApJ, 850, 52.
Rosenfeld, K. A., Andrews, S. M., Hughes, A. M., et al.
Lynden-Bell, D., & Pringle, J. E. 1974, MNRAS, 168, 603.
2013, ApJ, 774, 16.
Matr`a, L., Dent, W. R. F., Wyatt, M. C., et al. 2017,
Teague, R., Guilloteau, S., Semenov, D., et al. 2016, A&A,
MNRAS, 464, 1415.
592, A49.
010203040TB(K)(a)010203040TB(K)(b)−2000−1000010002000Velocity(ms−1)010203040TB(K)(c)ResampledDeprojected−π−π/20π/2πDiskPolarAngle−π−π/20π/2πDiskPolarAngle15
Teague, R., Bae, J., Bergin, E. A., et al. 2018, ApJ, 860,
Walsh, C., Daley, C., Facchini, S., et al. 2017, A&A, 607,
L12.
Teague, R., Henning, T., Guilloteau, S., et al. 2018, ApJ,
864, 133.
A114.
Weidenschilling, S. J. 1977, Ap&SS, 51, 153.
Teague, R. & Foreman-Mackey, D. 2018, Research Notes of
Williams, J. P. & Best, W. M. J. 2014, ApJ, 788, 59.
the American Astronomical Society, 3, 173.
Whipple, F. L. 1972, From Plasma to Planet, 211.
Teague, R. & Birnstiel, T. 2018, eddy: Extracting Disk
Dynamics v1.0, Zenodo, doi:10.5281/zenodo.1440051
van Boekel, R., Henning, T., Menu, J., et al. 2017, ApJ,
837, 132.
Yen, H.-W., Koch, P. M., Liu, H. B., et al. 2016, ApJ, 832,
204.
Yen, H.-W., Koch, P. M., Manara, C. F., et al. 2018, A&A,
Van Der Walt, S., Colbert, S. C., & Varoquaux, G. 2011,
616, A100.
Computing in Science & Engineering, 13, 22.
|
1202.5112 | 2 | 1202 | 2012-11-09T10:23:28 | An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397 | [
"astro-ph.EP",
"astro-ph.SR"
] | Searches for planetary transits carried out in open and globular clusters have yielded to date only a handful of weak, unconfirmed candidates. These results have been interpreted either as being insignificant, or as evidence that the cluster chemical or dynamical environment inhibits the planetary formation or survival. Most campaigns were limited by small sample statistics or systematics from ground-based photometry. In this work we performed a search for transiting planets and variables in a deep stellar field of NGC 6397 imaged by HST-ACS for 126 orbits. We analyzed 5,078 light curves, including a pure sample of 2,215 cluster-member M0-M9 dwarfs. The light curves have been corrected for systematic trends and inspected with several tools. No high-significance planetary candidate is detected. We compared this null detection with the most recent results from Kepler, showing that no conclusive evidence of lower planet incidence can be drawn. However, a very small photometric jitter is measured for early-M cluster members (<~2 mmag on 98% of them), which may be worth targeting in the near future with more optimized campaigns. Twelve variable stars are reported for the first time. | astro-ph.EP | astro-ph | Astronomy&Astrophysicsmanuscript no. paper
June 13, 2021
c(cid:13) ESO 2021
An HST search for planets in the lower main
sequence of the globular cluster NGC 6397⋆
V. Nascimbeni1,2,3⋆⋆, L. R. Bedin2,3, G. Piotto1,2, F. De Marchi4, and R. M. Rich5
2
1
0
2
v
o
N
9
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
2
1
1
5
.
2
0
2
1
:
v
i
X
r
a
1 Dipartimento di Astronomia, Universit`a degli Studi di Padova, Vicolo dell'Osservatorio 3, 35122 Padova, Italy
e-mail: [email protected], [email protected]
2 INAF -- Osservatorio Astronomico di Padova, vicolo dell'Osservatorio 5, 35122 Padova, Italy
e-mail: [email protected]
3 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218
4 Dipartimento di Fisica, Universit`a di Roma Tor Vergata and INFN, Sezione di Roma Tor Vergata, I-00133 Roma
e-mail: [email protected]
5 Division of Astronomy and Astrophysics, University of California, Los Angeles, 430 Portola Plaza, Box 951547,
Los Angeles, CA 90095-1547, USA. e-mail: [email protected]
Submitted December 16, 2011; accepted February 16, 2012
ABSTRACT
Searches for planetary transits carried out in open and globular clusters have to date yielded only a handful of weak, unconfirmed
candidates. These results have been interpreted as either being insignificant, or evidence that the cluster chemical or dynamical
environment inhibits planetary formation or survival. Most campaigns have been limited by small sample statistics or systematics
from ground-based photometry. We performed a search for transiting planets and variables in a deep stellar field of NGC 6397 imaged
by HST-ACS over 126 orbits. We analyzed 5,078 light curves, including a careful selection of 2,215 cluster-member M0 -- M9 dwarfs.
The light curves were corrected for systematic trends and inspected using several tools. No high-significance planetary candidate is
detected. We compared this null detection with the most recent results from Kepler, showing that no conclusive evidence of lower
planet incidence can be drawn. However, a very small photometric jitter is measured for early-M cluster members (. 2 mmag on 98%
of them), which may be worth targeting in the near future with more optimized campaigns. Twelve variable stars are reported for the
first time.
Key words. techniques: photometric -- stars: planetary systems -- clusters: individual: NGC 6397
1. Introduction
More than seven-hundred extrasolar planets are known (exo-
planet.eu database). Most of them are characterized only
through radial velocities (RV) and lack any information about
their "real" mass Mp, because only Mp sin i can be measured,
where i is the inclination of the orbit with respect to the line
of sight. Their size is also unknown, which prevents us from
getting any clues about their density and physical composition.
Exoplanetary transits are highly complementary to RV tech-
niques, providing the planetary radius Rp from the stellar radius
R⋆ in a direct geometrical way (Seager 2011, p. 55). Photometric
searches for transits can also go much deeper than RVs in mag-
nitude, and can monitor thousands of stars simultaneously. The
Kepler mission (Borucki et al. 2010), for instance, has demon-
strated the power of this technique by discovering many plan-
etary systems with unexpected properties. Hundreds of Kepler
"candidate planets", for which confirmation and mass measure-
ment via RVs remains infeasible, are still very useful for sta-
tistical purposes (Howard et al. 2011; Schlaufman & Laughlin
2011)
Star clusters, and in particular globular clusters (GC), offer
a unique opportunity to study how the chemical and dynamical
⋆ Based on observations with the NASA/ESA Hubble Space
Telescope, obtained at the Space Telescope Science Institute, which is
operated by AURA, Inc., under NASA contract NAS 5-26555.
⋆⋆ Visiting PhD Student at STScI (DDRF D0001.82432 program).
environments affect planetary formation and evolution. They are
also comprised of stars that share (in most cases) the same age
and chemistry, and whose radii R⋆ and masses M⋆ are reliably
known on their main sequence (MS). Open clusters (OC) have
been targeted for extensive transit searches (Mochejska et al.
2005; Montalto et al. 2007, 2011; Hartman et al. 2009, most
notably) but only a handful of weak, unconfirmed candidates
have been so far detected (Mochejska et al. 2006; Montalto et al.
2011). A global reanalysis suggests that the overall statisti-
cal significance of these campaigns is so low that it could be
compatible with the planet host incidence observed in the field
(van Saders & Gaudi 2011). On the other hand, GCs are on aver-
age much richer in stars than OCs, providing a much higher sta-
tistical significance in the case of a null detection. The only GCs
monitored for transits have been 47 Tucanae for both 8 days with
HST WFPC2 (Gilliland et al. 2000) and 33 days from the ground
(Weldrake et al. 2005); and ω Centauri with a 25-day ground-
based campaign (Weldrake et al. 2008). No planetary detection
has been claimed, and Gilliland et al. (2000) concluded that the
planetary occurrence in 47 Tuc is smaller by a factor of ten than
in field stars. The reasons that have mostly been hypothesized to
explain the lack of giant planets in GCs are their metallicity and
their dynamical environment.
It has long been known that metallicity is a strong primary
parameter that correlates with the fraction of stars with plan-
ets Φp. For giant planets, Fischer & Valenti (2005), among oth-
ers, measured an increase from the typical value Φ⊙p ∼ 0.03 for
1
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
stars with solar metallicity up to Φp ∼ 0.25 for very metal-
rich ([Fe/H] & +0.3) stars. For moderately metal-poor stars
(−0.5 . [Fe/H] . 0), it is still disputed whether this corre-
lation becomes flat at values Φp ≃ Φ⊙p (Udry & Santos 2007;
Santos et al. 2011) or Φp continues to decrease exponentially
towards lower metallicities (Johnson et al. 2010). As more and
more low-mass planets (Mp . 30M
) are being discovered,
⊕
it becomes clear that their Φp is much larger than that of gi-
ant planets (Lovis et al. 2009; Wittenmyer et al. 2011), probably
around ∼20 -- 30 %. The occurrence of low-mass planets seems
to be insensitive to the host star metallicity (Udry et al. 2006),
except maybe for very late-type stars (Johnson & Apps 2009;
Schlaufman & Laughlin 2011) but the M-dwarf metallicity cali-
bration is still too uncertain to draw conclusions.
The complex dynamical environment of a cluster is the sec-
ond major concern about the survival of protoplanetary sys-
tems. Theoretical studies give different answers, but some of
them imply that gravitational stripping processes are not strong
enough to disrupt very short-period (P < 5 days) planetary sys-
tems, even in the densest regions of a typical GC (Fregeau et al.
2006; Spurzem et al. 2009). Planets do indeed exist in clusters.
Giant planets have been detected around evolved stars belong-
ing to open clusters (Lovis & Mayor 2007; Sato et al. 2007), and
around a pulsar in the globular cluster M4 (Thorsett et al. 1999).
It is therefore difficult to understand why no planets have been
detected in 47 Tuc. The search by Gilliland et al. (2000) was sen-
sitive only to giant planets (Rp & 1Rjup) around stars in the upper
MS, excluding late-K and M dwarfs. More data are needed to
sample other regions of the parameter space.
On the other hand, it is convenient to search for transits
around KM main-sequence stars because of their larger expected
signal. These targets, even in the nearest GCs, are faint (V > 18)
and extremely crowded. Space-based observations are necessary
to achieve the needed signal-to-noise ratio (S/N) per time unit,
and to minimize the number of false positives caused by blended
photometric contaminants. The wide-field imagers mounted on
HST (ACS and WFC3) are unrivalled in this respect, and their
large archive of deep photometric series of GCs is already avail-
able to be exploited for transit searches. However, we show in
Sec. 3 that HST time-resolved photometry is influenced by many
sources of systematic errors that require a careful correction. We
developed specific tools to this purpose and applied them to a
test case. Other data sets (such as those for 47 Tuc) will be in-
vestigated in the near future, to test whether a more optimized
campaign is worth pursuing.
In this paper, we present a search for variables and plan-
etary transits in NGC 6397, exploiting a 126-orbit data set
from the HST Advanced Camera for Surveys (ACS; GO-
10424; PI:Richer). NGC 6397 is a metal-poor ([Fe/H] ≃ −2)
core-collapsed globular cluster, and the second-closest known
(Gratton et al. 2003; Richer et al. 2008). Our work is in some
sense complementary to that presented by Stello & Gilliland
(2009), who employed the same data set to study the microvari-
ability of highly saturated red giants. Though these data were not
optimized for a transit search, its unprecedented depth allowed
us to search for planets on a homogeneous sample of 2,215 mem-
ber M-dwarf stars down to the hydrogen-burning limit. The M
dwarfs are considered the most promising targets to discover
rocky planets in the near future (Scalo et al. 2007). They are the
), and therefore the transit depth
smallest stars (R⋆ ≃ 0.1 -- 0.5R
for a given planet is larger by a factor of ∼ 4 -- 100 that that of a
solar-type host star. They are also intrinsically faint (L⋆ ≃ 0.02 --
5·10−5L
), meaning that a habitable planet would have an orbital
⊙
period of only P = 10 -- 30 days.
⊙
2
2. Observations and data reduction
Our analysis is based on the HST data set GO-10424 (PI:
Richer), whose acquisition was originally made to probe the bot-
tom of the MS and the end of the white dwarf (WD) cooling se-
quence in NGC 6397 (Hansen et al. 2007; Richer et al. 2008). A
single 202′′ × 202′′ field, located 5′ from the cluster core was
imaged for 126 orbits with the wide-field channel of the ACS.
In each orbit, a single exposure through the F606W filter was
preceded and succeeded by two F814W images, with exposure
times ranging between 584 s and 804 s (median: 704 s). Overall,
252 F814W frames and 126 F606W frames were secured, for a
total exposure time of ∼ 50 hours. Shorter exposures (texp = 1, 5,
40 s) were also taken to measure the fluxes of the brightest stars,
but they were not used in our study.
The dynamic range of the "deep" exposures is perfect for our
purposes, as the observed luminosity function (LF) of the cluster
MS members peaks at mF814W ≃ 21 (that is, on well-measured
stars with S/N ∼ 200 in the same filter; Fig. 1, histogram on
the right panel). Saturation occurs around M0V spectral type
at mF814W ≃ 18.7 and the faintest cluster members (M9V) are
found at mF814W ≃ 24 (S/N ∼ 20), such that a sample of ∼ 2, 000
cluster M dwarfs down to the hydrogen-burning limit is available
to transit search (Fig. 1, left panel).
On the other hand, these data were not optimized to search
for transit events, thus two aspects of the observing setup are
somewhat restrictive. First, the time coverage of the frames
is discontinuous, unlike in Gilliland et al. (2000). The data for
∼ 50 h of integration time were spread among twenty-one non-
contiguous days, spanning a twenty-eight day period. This trans-
lated into a much lower completeness for our search, especially
for the longest periods (P > 3 d, see Sec. 5). The second reason
is that each pointing was shifted within a ten-position dither-
ing pattern ∆x, ∆y plus a subpixel offset δx, δy. While dithering
is usually suitable for undersampled images (Anderson & King
2000), it prevented us from reaching the highest photometric
accuracy possible for the brightest stars. For these targets, the
amount of random noise is so low that the unavoidable flat-
field and pixel-to-pixel residual errors are no longer negligible.
A zero-point correction, discussed in the next Section, was de-
veloped to suppress these systematic errors.
We carried out
the data reduction on individual bias-
corrected and flat-fielded .flt images provided by the HST
pipeline. We employed the master input list from Anderson et al.
(2008), and the code described by Anderson & King (2006)
based upon the effective PSF (ePSF) approach first developed by
Anderson & King (2000). Four tests were performed on a subset
of twenty F814W frames taken at the same integer-pixel dither-
ing position, to choose the best reduction strategy among:
1. Allowing spatially-constant perturbed ePSFs as described by
Anderson & King (2006);
2. Correcting the raw frames for charge transfer (in)efficiency
(CTE) with the pixel-based algorithm proposed by
Anderson & Bedin (2010);
3. Using both the 1) and 2) corrections;
4. Using neither.
For all the measured sources, the root mean square (RMS) values
of their light curve as a function of magnitude were compared.
Methods 1-3 above provided no detectable improvement over
the fourth choice, therefore we decided not to apply any CTE or
PSF correction at this stage. Actual PSFs are both spatially and
temporally variable, and require a frame-to-frame a posteriori
correction that is explained in Section 3.
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Fig. 1. Left panel. Color-magnitude diagram (mF606W − mF814W, mF814W) for all the stars in the Anderson et al. (2008) master list;
the 5,078 sources selected in this work lie between the saturation limit at mF814W ≃ 19 and the faint limit mF814W ≃ 26.5 (dash-dot
lines). The red line is the isochrone by Dotter et al. (2007) employed in Richer et al. (2008). The dotted line marks the loci occupied
by equal-mass MS-MS binaries. Right panel. The RMS for light curves in our sample as a function of the instrumental magnitude.
Red circles: median RMS averaged over 0.5 mag bins, without any correction. Cyan circles: the same, after PSF-ZP correction.
Green circles: both PSF-ZP and dither-ZP corrections applied. Small points are the RMS of the individual light curves after both
corrections. The solid orange line is the expected theoretical noise level. The superimposed histogram represents the number of
targets in each 0.5 mag bin (scale at the right).
Hereafter, we focus our analysis on only F814W frames for
many reasons. They are much deeper than F606W images for
faint, red, low MS stars, and less affected by PSF short-term vari-
ations as shown for this very same dataset by Anderson & King
(2006). The sky background is also much lower in F814W expo-
sures. The way in which the F814W frames were sampled (at the
beginning and the end of each 96-minute orbit) ensures that the
F606W frames are nearly useless for increasing the transit detec-
tion efficiency. In this way, we also avoided the need to correct
for any tricky registration of the light curves between the two
filters.
Sources beyond the saturation limit of the longest expo-
sures (mF814W . 19) and sources detected in fewer than 200
frames (mF814W & 26.5) were excluded from this study, leav-
ing 5,078 objects including cluster members, field stars, and a
limited number of non-stellar objects. We evaluated the mem-
bership of each entry by performing a selection on the proper
motions between our epoch and the archival ACS GO-11633
data set (PI: Rich), centered on the same field. We flagged 2,430
sources as cluster members, of which 215 belong to the white
dwarf (WD) sequence (see the CMD on Fig. 1, left panel). The
remaining 2,215 MS stars are red dwarfs ranging from M0V
) down to the hydrogen-
spectral type (Rp ≃ 0.5R
burning limit, as confirmed by superimposing an isochrone by
Dotter et al. (2007) on the CMD, as done in Richer et al. (2008).
, Mp ≃ 0.5M
⊙
⊙
3. Systematic correction
The instrumental magnitude −2.5 log(DN) of each star was reg-
istered to the median instrumental magnitude of stars measured
in the deepest frame of the F814W series. We refer to this mag-
nitude as m. Saturation occurs at m . −13.4. The RMS σm of
our full sample of 5,078 light curves was compared with the
expected noise budget, as calculated by combining theoretical
photon-, dark-, background- and readout noises (right panel of
Fig. 1, orange line). Individual σm were averaged whitin 0.5 mag
bins by applying a clipped median (red circles in the same plot)
to exclude outliers from the comparison.
On the bright side, the observed noise level on average is far
higher than expected, even by 50-60% for stars with m . −12
(red circles, right panel of Fig. 1). We identified the source
of most of this excess noise as a variation in the photomet-
ric zero point (ZP) induced by systematic changes in the PSF
shape. Long-term instability of the PSF was reported for ACS by
Anderson & King (2006). We noticed that this ZP change (here-
after called PSF-ZP) follows a well-defined pattern as a function
of time and average x, y position on the detector. The pattern can
be mapped by evaluating for each star two diagnostic parameters
that appear to be strongly correlated with the PSF-ZP shift:
1. The difference between the median magnitude hmibeg mea-
sured in frames taken at the beginning of the orbit, and the
median magnitude hmiend measured in frames taken at the
end of the orbit (left panel of Fig. 2, color-coded in the range
−0.02-0.02 mag from black to red).
2. The difference between the median magnitude of the star
hmi7th measured in 16 consecutive frames taken during the
seventh "visit" of the program (2453451 < JD < 245352),
and the median magnitude hmi of its full light curve (right
panel of Fig. 2, same scale).
3
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Fig. 2. Mapping the PSF-ZP shift as a function of time and position on the ACS detector, with two different diagnostics (see text).
The First (left panel) is the difference between the median magnitudes hmibeg and hmiend measured in frames taken respectively at
the beginning and the end of the orbit, while the second (right panel) is the difference between the median magnitude of the star
hmi7th measured during the seventh "visit" of the program (2453451 < JD < 245352), and the median magnitude hmi of its full light
curve. Top panel: All the high-S/N light curves (σm < 0.02) have been registered to their average magnitude hmi. In all panels, the
color scale spans the range −0.02-0.02 mag from black to red.
The pattern is very similar in both cases. The first diagnostic
hmibeg − hmiend is probably a proxy of the real origin of the PSF-
ZP systematic: a thermal/mechanical instability linked to the or-
bital phase. The dependence of PSF-ZP on time becomes evi-
dent when all the light curves with a high S/N (i.e., measured
on all the 252 frames and having σm < 0.02) are registered to
their average magnitude hmi, stacked in the same plot, and color-
coded as a function of hmi7th −hmi (Fig. 2, top panel, color scale
from black to red in the range −0.02 to 0.02 mag). It is clear
that on average stars whose flux is overestimated during the sev-
enth visit are also systematically underestimated in the last visits
(JD>2453460), and viceversa.
For a given frame and within the same chip, the PSF-ZP is a
smooth function of the position on the detector. The diagnostics
hmi7th − hmi and hmibeg − hmiend are too noisy when evaluated
for faint stars to implement an effective correction with them.
We chose instead to correct the PSF-ZP with a local approach,
adapted from the differential photometry algorithms described
in Nascimbeni et al. (2011). For each target star i in our sam-
ple, a set of N nearby reference stars k = 1, . . . , N was selected
with the following criteria: 1) they had to lie on the same chip
as the target and be within 200 pixels of it; 2) they had to be at
least 20 stars whose total flux had to exceed ten times the flux of
the target; and 3) they had to be detected in at least 250 frames
among 252, instead of the 200-frame limit required for a target
star. When requirement 2) was not met, the search radius was in-
j,k)/ Pk(1/σ2
creased until it was met. If on a given frame j a reference star k
was not detected, or when its magnitude m j,k was more than 3σ j,k
off its value averaged over the series hm j,ki, its magnitude was
set to hm j,ki. A reference magnitude m0, j was calculated in each
frame by performing the weighted mean of the magnitudes of the
N reference stars m0, j = Pk(m j,k/σ2
j,k) (Broeg et al.
2005). The target magnitude was then normalized to m0, j. The
PSF-ZP correction was applied only when it decreased the over-
all RMS of the target light curve. After the correction, the me-
dian RMS of the light curves (right panel of Fig. 1, cyan circles)
was substantially smaller. For the brightest stars (m . −12), it
had decreased to a level ∼ 15% above the expected noise.
Most of the remaining excess noise is due to local ZP
changes as well as the dithering pattern employed: an integer
∆x, ∆y shift plus a small sub-pixel offset δx, δy were added to
the initial pointing x0, y0 (where the units are in physical pixels).
We refer to this systematic effect as dith-ZP. To each light curve,
already corrected for PSF-ZP, we applied two decorrelating al-
gorithms where
1. The median magnitude of the star hmi∆x,∆y was calculated for
each subset of frames sharing the same integer-pixel dither
∆x, ∆y. The magnitude m of each frame in the subset ∆x, ∆y
was then registered to hmi∆x,∆y.
2. For each ∆x, ∆y subset corrected by the previous step, we
considered the magnitude m j in each frame j as a function of
4
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
the sub-pixel shift δx j, δy j. An ordinary linear least squares
fitting was carried out to find the coefficients c0, cx, and cy
such that
(1)
m j = c0 + cx δx j + cy δy j .
Once the best-fit value m′j = c0 + cx δx j + cy δy j had been
evaluated for each frame, the corrected light curves were ex-
tracted and normalized to zero by evaluating m′j − m j.
Both steps 1) and 2) were applied only when the RMS had
decreased. The resulting median RMS of the PSF-ZP + dith-ZP
correction is plotted in the right panel of Fig. 1 as green circles.
The small black points in the same plot show the RMS of each
single light curve of cluster members after our PSF-ZP and dith-
ZP corrections have been applied. Their noise level approaches
the theoretical limit in every magnitude bin, demonstrating that
the algorithms we applied were effective.
4. Light curve analysis
4.1. Searchfortransit-likeevents
We searched for transits in the full set of 5,078 light curves (that
is, including field stars) by applying the box-fitting least-square
algorithm (BLS, Kov´acs et al. 2002). For each curve, BLS was
applied to search for periodic dips of duration ∆ and depth δ
with 10,000 trial periods between 0.2 and 14 days. The relative
transit duration q = ∆/P was constrained to the values possible
for planetary transits around low-MS stars (R⋆ = 0.08 -- 1.4R
For each star, three detection diagnostics were calculated:
the signal residue (SR), the signal detection efficiency (SDE)
associated with the maximum peak in the BLS periodogram
(Kov´acs et al. 2002), and the detection S/N defined as
).
⊙
S/NBLS =
δ
σ · √nt
,
(2)
where σ is the (unbinned) photometric noise and nt the number
of data points sampled during transits.
Other more sophisticated detection diagnostics, such as the
"signal-to-pink" S/N (Pont et al. 2006; Hartman et al. 2008), are
robust if correlated noise σred ("red noise", Pont et al. 2006) is
present. However, they require knowledge of σred on timescales
close to ∆. This is difficult to evaluate in our data, as transits are
expected to be undersampled by the observing cadence. Transits
of a P ∼ 3 d planet around a M4V star are expected to last
R⋆P/(πa) ∼ 60 min at most, and only ∼ 30-40 min for later
types, while images are sampled every 32-64 min. However, the
amount of red noise here is very low, as demonstrated by the sim-
ilarity of the measured RMS to the theoretical one. Therefore, we
decided to employ both S/NBLS and SDE as detection diagnos-
tics.
To set a reliable detection criterium, 2,215 light curves were
simulated, that had the same sampling times ti and noise level as
the real M-dwarfs. A synthetic transit (following the analytical
model of Mandel & Agol 2002) of a 1 Rjup planet was injected
into each curve, with a random uniform distribution in P and
sin i where P was bounded to the range 1 -- 5 d, while sin i was
constrained to allow transits. The process was iterated 200 times
for a total of 450,000 injections. We then tried to recover the
transits with BLS, by setting the same parameters used for the
real search. A planet was defined as "recovered" if at least two
transits had been sampled, and if the estimated orbital period (or
a low-order harmonics: 2:1, 3:1, 3:2) matched the injected one.
The distributions of the "injected" and "recovered" transits in the
SDE vs. S/NBLS plane are plotted as black and red points in the
upper left panel of Fig. 3. By dividing the parameter space into
cells and evaluating the fraction of "recovered" over "injected"
transits (Fig. 3, lower left panel), we obtained an estimate of the
expected fraction f of real transits successfully detected ("true
positives"). We defined detection criteria that guarantee a frac-
tion of false positives that is smaller than 10% (that is, f ≥ 90%)
where
SDE ≥ 5.25
S/NBLS ≥ 65 − 9 · SDE
S/NBLS ≥ 7
(3)
as indicated by the blue line in Fig. 3. With this choice, the frac-
tion of false negatives (that is, real planets discarded by selection
criteria) is about 55%, and gets larger for planets smaller than
Jupiter. This is unavoidable if one wishes to keep the fraction of
false positives as low as possible.
The position of all the real sources in the (SDE, SNBLS) plane
is shown in the upper-right panel of Fig. 3. Only four stars among
the full sample meet the criteria set in Eq. (3) or get very close to
the threshold. These "borderline" targets (ID#0269, 1961, 5936,
and 7637) were inspected and cross-checked individually.
ID#269 light curve is crippled by a CCD bad pixel falling
just under the star in one of the dithering positions.
ID#1961 is contaminated by brighter surrounding stars, and
its flux drops off significantly in one fourth of the images. The
BLS signal is probably spurious.
, R⋆ = 0.32R
ID#5936 seems accurately measured by the reduction
pipeline, though it lies extremely close to a saturated star. Its
light curve should be treated with caution. The parameters of
the detected signal (P ∼ 2.1 d, δ = 0.08 mag, q = 0.025)
would be compatible with a ∼ 1 Rjup transiting body with zero
impact parameter, or with a grazing eclipsing binary. We clas-
sified ID#5936 as a cluster M-dwarf member by analyzing its
proper motion (with M⋆ = 0.34M
from its
⊙
color), though its position in the CMD diagram is offset from
the MS by 1.4 mag in mF814W and 0.2 mag in color (Fig 3, lower
right panel). This cannot be due to binarity alone, and maybe
the presence of a bright contaminant or the departure from the
normal evolution of the companion play a role. There is also a
non-negligible probability that ID#5936 is a field star having a
proper motion compatible with the common motion of the clus-
ter. It it worth noting that at least one data point with a similar
decrease in brightness (0.08 mag, that is ∼ 4σ) fell outside the
expected transit windows fitted by BLS, and that the estimated
duration (∆ = 76 min) is way larger than expected. We do not
consider ID#5936 to be a convincing planetary candidate. Its co-
ordinates are listed in Table 1 for possible further studies.
⊙
ID#7636 was rejected because it fell over a bad column on
frames corresponding to "transits".
In summary, we did not detect transits in our light curves,
at least with an acceptable degree of statistical significance. We
discuss the significance of this null detection in Section 5.
4.2. Searchforvariablestars
We performed a search for variable stars in our full database
of 5,078 light curves corrected for systematic errors. First, the
coefficient of spectral correlation (Ferraz-Mello 1981) was cal-
culated for each light curve. Following the method described in
de Marchi et al. (2007, 2010), we obtained a sample of 13 sus-
pected variable stars (Table 1). All these candidates, based on
their proper motions and position in the CMD, were identified as
5
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Fig. 3. Upper left: distribution of SDE and SNBLS from the BLS analysis of 450,000 artificially injected transits in synthetic light
curves. "Recovered" transits are plotted as red dots (see text). Lower left: same as above, where the parameter space has been
divided into cells and color-coded as a function of the fraction f of transits successfully recovered. Cells with f > 90% (that is,
with an incidence of false positives smaller than 10%) are highlighted with a black border. The blue line correponds to the threshold
defined in Eq. (3). Upper right: distribution of SDE and SNBLS from the BLS analysis of the full sample of 5,078 real light curves.
Four low-significance candidates are labeled. Lower right: location of the four low-significance candidates on the CMD.
field stars with high confidence. To classify these objects, a least
squares iterative sine-wave search was applied (Van´ıcek 1971).
Most of our candidates show a single harmonic sinusoidal
shape and short periods (P ≤ 9 d), namely ID#830, ID#7523,
ID#2178, ID#5600, and ID#4957. Without other elements, it
was impossible to derive an unambiguous classification for these
variables. We suspect that these stars are most probably field
BY Draconis variables, i.e. spotted and rotating KM dwarfs.
This tentative classification is supported by their very red col-
ors (mF606W − mF814W = 1.27-2.65).
For four stars, a good best-fit can be obtained using two har-
monics. The second harmonic in the light curves of ID#3428,
ID#1383, ID#4430, and ID#270 could indicate the presence of
spots on the surfaces and confirm the BY Dra-type classification,
while in the light curve ID#6119 the second harmonic reveals the
profile of a W UMa contact eclipsing binary system. Two stars
(ID#2086 and ID#258) show clearly orbit-to-orbit variability but
the time coverage was too short to infer reliable values for their
periods: we classified them as generic "long period variables".
Finally, the fluctuations in the light curve ID#1882 are too small
to allow us to confirm its nature as a variable star, hence we dis-
carded it from our analysis. The summary classification of the
entire sample is reported in Table 1, along with the best candi-
date transit found by BLS and discussed in Section 4.1.
5. Completeness and significance
The significance of our null detection of transits was assessed by
considering only the 2,215 cluster-member M dwarfs for which
we derived reliable estimates of R⋆ and M⋆ from their position
in the CMD.
For a planet of given radius Rp and orbital period P, the num-
ber of expected planet detections is given by
Np = N⋆ x (cid:16)Φp(P, Rp) · Φgeo · Φdet(cid:17) dP dRp ,
where Φp is the fraction of stars with a planet, Φgeo is the geo-
metric a priori probability for that system to be aligned such that
(4)
6
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Fig. 4. Top panels: light curves of the variable stars found (first twelve entries in Table 1), folded around the best-fit period. Bottom
left panel: position of the variables (red IDs) in the (mF606W − mF814W,mF814W) color-magnitude diagram. Bottom right panel:
photometric RMS of the variables (red IDs) compared with all the analyzed light curves.
a transit occurs, Φdet is the probability for that transit to be de-
tected by our pipeline, and N⋆ is the number of target stars. As in
our case we found that Np = 0, we wished to estimate an upper
limit to Φp, at least in the (P, Rp) range for which the efficiency
of our search Φdet is not negligible. Φdet is expected to depend on
the transit depth (Rp/R⋆)2, the duration ∆, and the orbital period
P.
For simplicity, our analysis was limited to two values of
planetary radii: "Jupiter" planets (1 Rjup) and "Neptune" planets
(0.338 Rjup). To estimate Φdet, we ran simulations in a way simi-
lar to what has been done to set the detection threshold (Section
4.1). In each set of 2,215 simulated light curves, a synthetic tran-
sit was injected into each curve, with random uniform distribu-
tions of both P and sin i (1 ≤ P ≤ 5 d, sin i was constrained to
allow transits). The process was iterated 200 times, for a total of
450,000 injections. We then tried to recover transits with BLS,
by setting the same parameters and detection criterium defined
in Eq. (3) and adopted for the real search. To derive Φdet, the
ratio of the detected to injected transits was evaluated for each
0.1-day bin of period P. The resulting distribution is plotted as a
red line in the left panels of Fig. 5, as a function of the injected
orbital period Pin. For comparison, the ratio of the "recoverable"
transits (i.e., with at least two transits sampled and P/Pin =1:1,
3:2, 2:1) to the injected transits is plotted with a green line on
the same panels. As expected, Φdet is a decreasing function of
P, with minor features at integer and semi-integer values of P
owing to phasing effects. We note that for "Neptunes", Φdet is
extremely low (0.005 -- 0.01) even for short periods (P ∼ 1 -- 2 d).
7
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Table 1. Parameters for the variable stars found.
ID#
RA (2000.0)
h:m:s
17:41:06.078
17:41:11.653
17:41:14.751
17:40:58.818
17:41:03.155
17:41:13.183
17:40:55.370
17:41:09.636
17:41:00.192
17:41:01.787
17:41:09.888
17:41:14.789
17:41:10.379
17:40:59.316
3428
1383
270
6119
4430
830
7523
2178
5600
4957
2086
258
1882
5936
d:m:s
DEC (2000.0) mF814W mF606W − mF814W
−53:45:47.62
−53:45:02.54
−53:45:04.94
−53:45:41.10
−53:44:49.11
−53:45:37.72
−53:45:27.03
−53:43:17.61
−53:42:46.28
−53:43:31.37
−53:43:10.59
−53:44:58.36
−53:44:52.05
−53:44:06.51
21.884
20.538
20.110
20.054
20.456
22.598
20.306
19.362
19.979
20.778
19.359
19.786
19.430
20.524
1.513
1.848
1.213
1.367
1.761
2.654
2.463
2.146
1.271
2.520
1.097
0.813
0.792
1.493
p.m.(α)
(pixels)
−1.2192
−1.3037
−1.0478
−1.4436
−1.2356
−1.3962
1.0304
−1.6238
−1.5276
−1.6460
−1.5737
−1.7434
−1.0721
0.0181
p.m.(δ)
(pixels)
0.4788
0.2494
0.7856
0.5940
0.4401
0.6337
0.5997
0.5057
0.3267
0.3394
0.3106
0.3614
0.7363
0.0055
P
(days)
2.860
2.650
2.290
0.378
1.095
2.700
2.520
2.790
≈ 8.9
≈ 8
≈ 12
≈ 25
0.920
2.120
∆m
(mag)
0.047
0.025
0.044
0.051
0.069
0.026
0.009
0.007
0.025
0.019
0.006
0.008
0.005
0.080
Nharm
notes
2
2
2
2
2
1
1
1
1
1
1
1
1
--
double-wave BY Dra
double-wave BY Dra
double-wave BY Dra
W UMa
BY Dra?
BY Dra?
BY Dra? (weak)
BY Dra? (weak)
BY Dra?
BY Dra?
long period
long period
spurious candidate?
grazing binary?
Notes. The columns give: the ID number of the star, the right ascension α and declination δ at epoch 2000.0, the calibrated magnitude m(F814W)
and color m(F814W)−m(F606W), the proper motion (α,δ) in ACS-WFC pixels relative to the cluster, the period found in days, the amplitude
found in magnitude, and the number of harmonics employed in the fit and a tentative interpretation.
This is a consequence of inefficient sampling, which makes the
significance of neptunian transits very weak: 3 -- 4σ even for the
most favourable case.
The geometric factor Φgeo was calculated for each injected
transit as (Rp + R⋆)/a (a is the semimajor axis), and then con-
volved with Φdet to obtain the probability of detecting transits
on a star which is known to host a planet on a random orbit, as a
function of its period (Fig. 5, middle panels).
We parametrized the "planet occurrence" following the anal-
ysis of Howard et al. (2011) for the distribution of 1,235 plane-
tary candidates detected by Kepler. In this case, Φp(P) was as-
sumed to be a power law modified with an exponential cut-off at
period Pcut
dΦp(P)
d log P
= k · P β (cid:16)1 − e−(P/Pcut)γ(cid:17) .
(5)
⊕
⊕
⊕ < Rp < 32R
From Howard et al. (2011) we adopted the parameters: k =
0.0025, β = 0.37, Pcut = 1.7 days, γ = 4.1 for "Jupiters"
(8R
), and k = 0.002, β = 0.79, Pcut = 2.2
days, and γ = 4.0 for "Neptunes" (4R
). Φp(P) is
plotted, with an arbitrary normalization, as blue squares in the
middle panels of Fig. 5.
⊕ < Rp < 8R
We first normalized Φp(P) by imposing P5
1 Φp(P) = 1, that
is assuming one planet with 1 d < P < 5 d per star. The total
number of expected detections Nexp(P Φp = 1) within each bin
over the range 1 < P < 5 d is (Φdet · Φgeo · Φp) N⋆. By summing
over the range 1 < P < 5 d, we obtained Nexp = 23.8 expected
detections of "Jupiters" and 0.14 of "Neptunes" (right panels of
Fig. 5, blue symbols). For a flat Φp(P) distribution, Nexp is larger
at 42.3 and 0.45, respectively.
The upper limit Φp,max to P5
1 Φp(P) suggested by our null
detection can be evaluated by simple binomial statistics, nor-
malizing Φp in order to get a 68.27% (1σ) or 95.44% (2σ)
probability of zero detections. We estimated for Jupiters that
Φp,max(1σ) = 4.8% and Φp,max(2σ) = 12.9% assuming the
Howard et al. (2011) Φp, and Φp,max(1σ) = 3.3%, Φp,max(2σ) =
9.1% assuming a flat Φp. As expected, Φp,max is well above unity
for the "Neptune" sample, leaving this planetary population es-
sentially unconstrained by our data (Fig. 5, lower right panel).
6. Discussion and conclusions
We have performed a search for planetary transits and variabil-
ity among 5,078 stars imaged in one of the deepest ACS fields
ever observed, which had been originally acquired to probe the
bottom of the main sequence of the metal-poor globular cluster
NGC 6397. The sample includes 2,215 M0-M9 dwarfs of secure
membership. Though these data were not optimized for such a
study, this is the largest homogeneous sample of M dwarfs ever
searched for variability.
Instrumental drifts and systematic errors caused by dithering
required a careful empirical correction, described in Section 3.
We developed and implemented algorithms that allowed us to
approach the theoretical noise limit across the whole magnitude
range 19 . mF814W . 26. The brightest cluster members (M0V)
were measured with an average scatter of 0.003-0.004 mag over
a time span of 28 days, illustrating the power of our decorrelating
techniques and the feasibility of transit searches in the low main
sequence of GCs.
We found no valid planetary transit above the significance
threshold that we set from simulations. Considering only cluster
stars, whose physical parameters can be reliabily infered, this
null detection sets an upper limit to the fraction of stars host-
ing a P < 5 d Jupiter-sized planet to about Φp = 4.8% at
1-σ confidence and 12.9% at 2-σ, assuming the planetary ra-
dius distribution derived by Howard et al. (2011) from Kepler
data. In other words, only 0.13 detections are expected assum-
ing that the underlying planetary population is similar to that
studied by Kepler. Most studies based on RVs also hypotesized
Φp < 1% for short-period, Jupiter-sized planets around solar-
type stars (Marcy et al. 2005). Furthermore, Φp is expected to be
a very steep function of the stellar metallicity (Fischer & Valenti
2005). Therefore, we are unable to make any firm conclusion
about the occurrence of giant planets in NGC 6397.
As demonstrated in Section 4.1, our data set is not sensitive
enough to Neptune-sized planets to draw any conclusion about
their occurrence, though a much higher Φp is expected for M
dwarfs by Howard et al. (2011) and Lovis et al. (2009), among
others. This was due to an inefficient sampling of the time series
available from archive material, which translated into poor phase
8
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Fig. 5. Completeness tests for cluster stars based on artificially injected transits, for 1 Rjup planets (upper row) and 0.336 Rjup planets
(lower row). Left panels: detection efficiency Φdet as a function of the input period, for planets potentially recoverable (green
symbols) and for those planets effectively detected by the criterion in Eq. (3) (red symbols). Middle panels: same as above, but
Φdet is convolved with the geometrical probability Φgeo for a planet to transit. The Howard et al. (2011) Φp(P) period distribution
function is plotted in blue symbols (arbitrary normalization). Right panels: number of expected transit detections per period bin,
assuming one planet per star within 1 < P < 5 d. Φp(P) is assumed to be flat (red symbols) or as modeled by Howard et al. (2011)
(blue symbols). 1- and 2-σ upper limits for the planet occurence Φp,max are shown.
coverage and severe undersampling of transit-like events, whose
duration is expected to be on the same order as the effective ca-
dence.
Twelve new variable stars have been identified in the
NGC 6397 field (Table 1). Most of these can be classi-
fied as BY Draconis variables, that is, spotted rotating KM
dwarfs. Interestingly, no variable has been detected among the
2,430 cluster members. Hundreds of member early-M dwarfs
(mF814W < 21, Fig. 1) follow the expected noise on timescales
of up to 28 days, though they were measured with a 0.003-0.006
mag precision. The lack of eclipsing binaries is unsurprising.
The number of expected detections can be estimated by scaling
down the number of EBs detected by Albrow et al. (2001) for
47 Tuc. If one takes into account the smaller fraction of binaries
in NGC 6397 (< 3%, even at radii smaller than the half-mass
radius, Milone et al. 2011) and the smaller number of targets
(2,215 vs. 46,422), we would expect much less than one EB. On
the other hand, the lack of BY Dra variables is more puzzling, as
one would expect 4-5 such detections, considering the number
of monitored cluster stars in this paper. We note that our stars
are cooler than the Albrow et al. (2001) sample, and we should
expect longer photometric periods. Our search is insensitive to
periods & 28 days, so this could be a possible explanation.
jit = σ2
obs − σ2
We evaluated an upper limit to the photometric jitter σjit of
the brightest members (M0V) by subtracting the contribution of
the expected noise σexp to the measured scatter σobs (that is, as-
suming σ2
exp). The fraction of stars f with σjit > 2
mmag is f . 2%. This value should be compared with the re-
sults found by Ciardi et al. (2011) examining the first quarter of
Kepler photometry on 2,182 field M dwarfs: these data cover
an interval of 33 days with an average cadence of 30 minutes,
which is quite similar to our cadence and time scale. They found
a fraction f ≃ 20% of stars with σ > 2 mmag, that is at least
an order of magnitude larger fraction than that we measured
in NGC 6397. The low MS of this cluster is extremely stable
and therefore worth targeting using more optimized observa-
tions. The James Webb Space Telescope Near-Infrared Camera
(NIRCam), for instance, would be able to probe the bottom of
the MS of NGC 6397 (mF814W ∼ 24) with a photometric preci-
sion better than 0.01 mag in a single 600 s exposure, without any
of the coverage/sampling issues mentioned above.
9
Nascimbeni et al.: An HST search for planets in the lower Main Sequence of the globular cluster NGC 6397
Anderson, J. & Bedin, L. R. 2010, PASP, 122, 1035
Anderson, J. & King, I. R. 2000, PASP, 112, 1360
Anderson, J. & King, I. R. 2006, PSFs, Photometry, and Astronomy for the
ACS/WFC, Tech. rep.
Anderson, J., King, I. R., Richer, H. B., et al. 2008, AJ, 135, 2114
Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977
Broeg, C., Fern´andez, M., & Neuhauser, R. 2005, Astronomische Nachrichten,
326, 134
Ciardi, D. R., von Braun, K., Bryden, G., et al. 2011, AJ, 141, 108
de Marchi, F., Poretti, E., Montalto, M., Desidera, S., & Piotto, G. 2010, A&A,
509, A17
de Marchi, F., Poretti, E., Montalto, M., et al. 2007, A&A, 471, 515
Dotter, A., Chaboyer, B., Jevremovi´c, D., et al. 2007, AJ, 134, 376
Ferraz-Mello, S. 1981, AJ, 86, 619
Fischer, D. A. & Valenti, J. 2005, ApJ, 622, 1102
Fregeau, J. M., Chatterjee, S., & Rasio, F. A. 2006, ApJ, 640, 1086
Gilliland, R. L., Brown, T. M., Guhathakurta, P., et al. 2000, ApJ, 545, L47
Gratton, R. G., Bragaglia, A., Carretta, E., et al. 2003, A&A, 408, 529
Hansen, B. M. S., Anderson, J., Brewer, J., et al. 2007, ApJ, 671, 380
Hartman, J. D., Gaudi, B. S., Holman, M. J., et al. 2008, ApJ, 675, 1254
Hartman, J. D., Gaudi, B. S., Holman, M. J., et al. 2009, ApJ, 695, 336
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2011, ArXiv e-prints
Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R. 2010, PASP, 122,
905
Johnson, J. A. & Apps, K. 2009, ApJ, 699, 933
Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369
Lovis, C. & Mayor, M. 2007, A&A, 472, 657
Lovis, C., Mayor, M., Bouchy, F., et al. 2009, in IAU Symposium, Vol. 253, IAU
Symposium, 502 -- 505
Mandel, K. & Agol, E. 2002, ApJ, 580, L171
Marcy, G., Butler, R. P., Fischer, D., et al. 2005, Progress of Theoretical Physics
Supplement, 158, 24
Milone, A. P., Piotto, G., Bedin, L. R., et al. 2011, arxiv:1111.0552
Mochejska, B. J., Stanek, K. Z., Sasselov, D. D., et al. 2006, AJ, 131, 1090
Mochejska, B. J., Stanek, K. Z., Sasselov, D. D., et al. 2005, AJ, 129, 2856
Montalto, M., Piotto, G., Desidera, S., et al. 2007, A&A, 470, 1137
Montalto, M., Villanova, S., Koppenhoefer, J., et al. 2011, A&A, 535, A39
Nascimbeni, V., Piotto, G., Bedin, L. R., & Damasso, M. 2011, A&A, 527, A85+
Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231
Richer, H. B., Dotter, A., Hurley, J., et al. 2008, AJ, 135, 2141
Santos, N. C., Mayor, M., Bonfils, X., et al. 2011, A&A, 526, A112+
Sato, B., Izumiura, H., Toyota, E., et al. 2007, ApJ, 661, 527
Scalo, J., Kaltenegger, L., Segura, A. G., et al. 2007, Astrobiology, 7, 85
Schlaufman, K. C. & Laughlin, G. 2011, ArXiv e-prints
Seager, S. 2011, Exoplanets, ed. Piper, S.
Spurzem, R., Giersz, M., Heggie, D. C., & Lin, D. N. C. 2009, ApJ, 697, 458
Stello, D. & Gilliland, R. L. 2009, ApJ, 700, 949
Thorsett, S. E., Arzoumanian, Z., Camilo, F., & Lyne, A. G. 1999, ApJ, 523, 763
Udry, S., Mayor, M., Benz, W., et al. 2006, A&A, 447, 361
Udry, S. & Santos, N. C. 2007, ARA&A, 45, 397
van Saders, J. L. & Gaudi, B. S. 2011, ApJ, 729, 63
Van´ıcek, P. 1971, Ap&SS, 12, 10
Weldrake, D. T. F., Sackett, P. D., & Bridges, T. J. 2008, ApJ, 674, 1117
Weldrake, D. T. F., Sackett, P. D., Bridges, T. J., & Freeman, K. C. 2005, ApJ,
620, 1043
Wittenmyer, R. A., Tinney, C. G., Butler, R. P., et al. 2011, ArXiv e-prints
Fig. 6. (not in the published version). Upper panel: relative
proper motions for all individual stars (black dots) with the
twelve identified variables plotted with red circles. The blue cir-
cle marks the average proper motion of NGC 6397. Lower panel:
mean astrometric error as a function of instrumental magnitude,
color-coded as above. See text in the final note for details. In
both panels, units are ACS physical pixels.
Most of the analysis techniques presented in this paper can
also be applied (with little or no modification) to other exist-
ing ACS/WFC3 time series of rich stellar fields. This is the case
for the metal-rich globular cluster 47 Tucanae, which has been
imaged with ACS and WFC3 over a longer time-frame. Such
a search for transits will allow us to complement the results of
Gilliland et al. (2000) in a different range of spectral types and
planetary masses.
Acknowledgements. This work was partially supported by PRIN INAF 2008
"Environmental effects in the formation and evolution of extrasolar planetary
system". V.N. acknowledges support by STScI grant DDRF D0001.82432. We
thank Ennio Poretti for helping us to identify the variable stars. We thank Aaron
Dotter for providing us with the isochrones used in Richer et al. (2008) and pre-
sented in Dotter et al. (2007). Some tasks of our data analysis have been carried
out with the VARTOOLS code (Hartman et al. 2008). We thank Ron Gilliland
for his useful comments and suggestions.
NOTE to the astro-ph version. The referee asked us if it is
possible that the proper motions have systematic problems for
the variable stars.
We doubt that variability could have biased our proper mo-
tions. In fact, our variables have proper motions statistically
consistent with those measured for non-variable field stars (see
Fig. 6, upper panel, red points). Also the astrometric error is con-
sistent with that found on stars of similar magnitude, with the
only exception of ID#4430 (Fig. 6, lower panel). ID#4430 is the
only variable contaminated by a nearby star; however, it does not
lie close to the MS in the CMD.
References
Albrow, M. D., Gilliland, R. L., Brown, T. M., et al. 2001, ApJ, 559, 1060
10
|
1905.03802 | 1 | 1905 | 2019-05-09T18:06:34 | Titan's Dynamic Love Number Implies Stably-Stratified Ocean | [
"astro-ph.EP"
] | The dynamic quadrupole Love number of Titan measured by \Cassini is $k_\mathrm{2,obs}=0.616\pm 0.067$, strongly indicating a global subsurface ocean. However, the theoretical Love number due to equilibrium tides is at most $k_\mathrm{2,eq}^\mathrm{max}\approx 0.48$ in the absence of an ice shell on top of the ocean. In reality, there is an outer ice shell of thickness $ 100\,\mathrm{km}$, reducing the equilibrium-tide Love number to $k_\mathrm{2,eq}\approx 0.42$. Therefore, other types of tidal response, like dynamic tides, may be also present in Titan. We propose that the ocean is stably stratified. As a result, there exist standing ocean waves (gravity modes) with eigen-frequencies close to the tidal frequency. Such a gravity mode (g-mode) is resonantly excited. It bends the outer ice shell radially and thus enhances the dynamic Love number by $k_\mathrm{2,g}$. In order for $k_\mathrm{2,g}$ to account for the discrepancy between $k_\mathrm{2,eq}$ and $k_\mathrm{2,obs}$, the Brunt-Vaisala frequency in the ocean is required to be $3.3\times 10^{-4}\,\mathrm{rad\, s^{-1}}$. It is compatible with the volatile-rich model for Titan that was proposed to explain the methane-rich atmosphere. The three components of the tidal potential with azimuthal degrees, $m=-2,0,2$, correspond to the three components of the quadrupole Love number, $k_\mathrm{2,-2}$, $k_\mathrm{2,0}$ and $k_\mathrm{2,2}$. They can excite retrograde, axisymmetric and prograde g-modes equally in the absence of rotation. However, Coriolis force induced by Titan's rotation breaks the symmetry among these modes. Most likely, only one of the Love-number components is significantly enhanced by a g-mode, while the other two are still attributed to equilibrium tides. This prediction is testable by observation. If confirmed, the smaller components of the Love number can be used to constrain the thickness of the outer ice shell. | astro-ph.EP | astro-ph |
Titan's Dynamic Love Number Implies
Stably-Stratified Ocean
Jing Luan a,
aInstitute for Advanced Study, Princeton, NJ 08540 (U.S.A.)
Copyright c(cid:13) 2019 Jing Luan
Number of pages: 25
Number of tables: 1
Number of figures: 1
Preprint submitted to Icarus
13 May 2019
Proposed Running Head:
Stably Stratified Ocean in Titan
Please send Editorial Correspondence to:
Jing Luan
Institute for Advanced Study
BH-128, 1 Einstein Drive
Princeton, NJ 08540, USA.
Email: [email protected]
Phone: (626) 437-6258
2
ABSTRACT
oretical Love number due to equilibrium tides is at most kmax
The dynamic quadrupole Love number of Titan measured by Cassini is k2,obs =
0.616± 0.067, strongly indicating a global subsurface ocean. However, the the-
2,eq ≈ 0.48 in the
absence of an ice shell on top of the ocean. In reality, there is an outer ice
shell of thickness ∼ 100 km, reducing the equilibrium-tide Love number to
k2,eq ≈ 0.42. Therefore, other types of tidal response, like dynamic tides, may
be also present in Titan.
We propose that the ocean is stably stratified. As a result, there exist stand-
ing ocean waves (gravity modes) with eigen-frequencies close to the tidal fre-
quency. Such a gravity mode (g-mode) is resonantly excited. It bends the
outer ice shell radially and thus enhances the dynamic Love number by k2,g.
In order for k2,g to account for the discrepancy between k2,eq and k2,obs, the
Brunt-Vaisala frequency in the ocean is required to be 3.3× 10−4 rad s−1. It is
compatible with the volatile-rich model for Titan that was proposed to explain
the methane-rich atmosphere.
The three components of the tidal potential with azimuthal degrees, m =
−2, 0, 2, correspond to the three components of the quadrupole Love number,
k2,−2, k2,0 and k2,2. They can excite retrograde, axisymmetric and prograde g-
modes equally in the absence of rotation. However, Coriolis force induced by
Titan's rotation breaks the symmetry among these modes. Most likely, only
one of the Love-number components is significantly enhanced by a g-mode,
while the other two are still attributed to equilibrium tides. This prediction
is testable by observation. If confirmed, the smaller components of the Love
number can be used to constrain the thickness of the outer ice shell.
3
Keywords: Titan; Saturn; Satellites, dynamics
4
1
Introduction
Titan, the most massive satellite of Saturn, is on a 15.9-day orbit with eccen-
tricity, e ≈ 0.0288, and inclination, I ≈ 0.35◦, with respect to the equatorial
plane of Saturn (Seidelmann, 1992). Titan's spin frequency, ΩT, is synchro-
nized with its orbital mean motion, norb. Saturn exerts a tidal potential on Ti-
tan. Its leading order is quadrupole, Φtide. It contains a static or time-averaged
part, Φstatic. Φtide reduces to Φstatic for a synchronously rotating satellite on
a circular orbit. The tidal deformation induced by Φstatic has relaxed to that
of a fluid body. The tidal deformation yields a gravitational potential that is
proportional to Φstatic by the dimensionless fluid Love number, kf. Titan's kf is
about 1.01, measured by Iess et al. (2010) and recently confirmed by Durante
et al. (2019). The fluid Love number for a homogeneous body is 1.5. Mass
concentrating towards the center results in kf < 1.5. The fluid Love number
and hydrostatic gravitational parameters (Durante et al., 2019) constrain the
mass distribution in Titan (e.g. Rappaport et al., 2008; Baland et al., 2014).
For a synchronously rotating satellite on an eccentric orbit, the distance and
the longitude of Saturn in the rest frame of the satellite both oscillate by
a fraction ∼ e at frequency norb. Consequently, Φtide contains a time-varying
part, Φe. The leading order of Φe is ∝ e and oscillates with frequency norb (Ap-
pendix B). The corresponding tidal deformation depends on material rigidity
of the solid parts of Titan, because the elastic response cannot be relaxed
within one orbital period. The induced gravitational potential is proportional
to Φe by the dynamic quadrupole Love number, k2. We mainly deal with k2
and call it the dynamic Love number. We refer to kf as the fluid Love number.
For a coherent solid body, k2 (cid:28) kf. A global liquid layer raises k2 towards
5
kf, but still k2 < kf unless the whole body is liquid. The gravitational po-
tential of Titan was measured by tracking Doppler shifted radio signals from
Cassini during its flybys of Titan. The static tidal response is associated with
kfΦstatic. The dynamic tidal response is associated with k2Φe ∼ k2eΦstatic. Be-
cause e (cid:28) 1 and k2 < kf, it is more difficult to measure k2 than kf. Iess et al.
(2012) measured Titan's dynamic quadrupole Love number and recently Du-
rante et al. (2019) improved it to k2,obs = 0.616 ± 0.067 (1σ). 1 The k2,obs is so
close to kf that it strongly indicates an internal ocean overlaid by a thin outer
ice shell.
The tidal response of Titan is usually treated as equilibrium tides because the
dynamic frequency ωdyn = (GMT/R3)1/2 ≈ 7.25×10−4 rad s−1, is much greater
than the tidal frequency, norb ≈ 4.56×10−6 rad s−1. The dynamic Love number
attributed to equilibrium tides is k2,eq. It increases with a thinner outer ice
shell, because the shell resists tidal deformation. It reaches the maximum value
2,eq ≈ 0.48 in the absence of an outer ice shell, but kmax
kmax
1σ below k2,obs. The theoretical kmax
2,eq mainly depends on the mass and size of
2,eq is still more than
the solid core and the rigidity of the core. The former is well constrained by
the observed static gravitational potential of Titan (Iess et al., 2010; Durante
et al., 2019). The latter may be artificially reduced in order to increase the
theoretical kmax
2,eq . However, as already pointed out in Rappaport et al. (2008)
and Durante et al. (2019), one has to adopt some yet unknown peculiar elastic-
viscous model for the core or to reconcile the tension between kmax
2,eq and k2,obs.
1 Durante et al. (2019) also tried to fit the time lag between Φe and k2Φe. It is
consistent with zero and its upper limit is about a tenth of the orbital period.
2 ≈ 0.082 ± 0.118, which is poorly constrained by
The imaginary part of k2 is kim
observation. We only discuss the real part of k2 in this paper.
6
Another way to reconcile the tension is to assume dense water (Baland et al.,
2014). We constrain our discussion to normal values for physical quantities as
shown in Table 1.
In reality, the outer ice shell has a finite thickness (d) and thus makes k2,eq
smaller than kmax
2,eq . Considering the balance between the radiogenic heating of
the rocky core and the conductive heat loss rate through the outer ice shell,
we estimate d ∼ 100 km. 2 3 With d = 100 km, k2,eq ≈ 0.42, according to the
tidal model described in Appendix A. This value is about 3σ below k2,obs. One
might ask whether tidal heating could generate enough power to maintain a
shell much thinner than 100 km and thus boost up k2,eq to almost its maximum
value. We checked several mechanisms for tidal heating but none of them can
do so (Appendix D).
Therefore, we are in need of a mechanism that could enhance the dynamic
2 We adopt the specific radiogenic heating rate of the Earth's mantle, 7.4 ×
10−8 erg s−1 g−1. The intrinsic heating of the Earth is partly due to radiogenic heat-
ing and primordial heating. The two resources are believed to be comparable. Given
that Titan's rocky core is of half of its total mass, Mc ∼ 0.5MT ∼ 6.7 × 1025 g, we
obtain a heating rate ∼ 5 × 1018 erg s−1. It could balance the conductive heat loss
Eloss ∼ 4πR2kice∆T /d ∼ 3.0 × 1018(d/100 km)−1 erg s−1, at ice shell thickness
rate,
d ∼ 100 km.
3 The obliquity of Titan, the angle between the spin axis and the orbital normal,
was measured to be 0.3◦ (Meriggiola and Iess, 2012). Because it is bigger than 0.12◦
expected for a completely solid Titan in a Cassini state, Bills and Nimmo (2011)
suggested that there existed a subsurface ocean. Unfortunately, the obliquity cannot
put a tight constraint on d. According to Baland et al. (2011), d could range from
a few tens of kilometers to two hundred kilometers.
7
Love number by (k2,obs − k2,eq)/k2,eq ≈ 46%. Since k2 is mainly attributed to
the radial displacement of Titan's surface, 4 increasing k2 requires enlarging
the surface's radial displacement. We propose that resonantly excited grav-
ity modes (g-mode) in the presumably stably-stratified ocean could account
for the enhancement. A g-mode is a standing ocean wave which contains
multiple half-wavelengths in the radial direction (Unno et al., 1989). The
number of nodes along the radial direction, nmode, is known as the radial
order. Stable stratification slows down the group velocity of ocean waves since
vg = ∂ω/∂kr ∼ ω2/(N k⊥), where k⊥ =
Λ/r is the horizontal wavenumber
√
and Λ = (cid:96)((cid:96) + 1) = 6 for quadrupole modes (Unno et al., 1989). Here N > 0
is the Brunt-Vaisala frequency. A larger N means the ocean is more stably-
stratified and thus leads to a slower group velocity. In general, a stabilizing
composition gradient or temperature gradient can both lead to a positive N .
The Earth ocean is stably stratified mainly by a stabilizing temperature gra-
dient. Thanks to the slow group velocity, g-modes have eigen-frequencies well
below the dynamic frequency, ωdyn. Thus it is possible to find a high-order
g-mode whose eigen-frequency matches the tidal frequency, norb. Such a mode
could be resonantly excited to a large enough amplitude to account for the
4 At Titan's surface, the density jumps from that of ice to almost zero, therefore
the radial displacement of Titan's surface leads to column mass excess (or deficit)
of order ρi1ξsurf
r
. This column mass excess (deficit) yields a perturbation to the
gravitational potential of Titan. In Titan's interior, e.g. at the interface between
the rocky core and the high-pressure ice layer, the density changes by ρc − ρhp. The
radial displacement at this interface also leads to a perturbation to the gravitational
potential of Titan. All these perturbations add up to the total induced tidal poten-
tial k2Φe. Because the density jump at Titan's surface is the biggest, k2 is mainly
attributed to the radial displacement at Titan's surface.
8
enhancement in the dynamic Love number.
A g-mode acts like a damped harmonic oscillator driven near resonance with
the frequency mismatch much bigger than the linear damping rate. Its ampli-
tude is inversely proportional to the frequency mismatch between the eigen-
frequency and the driving (tidal) frequency, δω ≡ ω − norb. The tidal fre-
quency generally falls between the eigen-frequencies of two g-modes with
neighboring radial orders. Therefore, δω is generally less than half of the fre-
quency separation, ∆ω, between the two modes. The frequency separation
is mainly determined by the Brunt-Vaisala frequency. Assuming the general
case, δω ∼ 0.5∆ω, we estimate Nreq ∼ 3.3 × 10−4 rad s−1 in order to enhance
the theoretical k2 to match what's observed. The required value for N is com-
patible with the volatile-rich model for Titan (Lunine and Stevenson, 1987).
In this model, the outer ice shell is enriched with methane, while the ice layer
beneath the ocean is enriched with ammonia. Since the molecular weight of
ammonia is greater than that of methane, there may exist a stabilizing com-
position gradient in the ocean yielding N ∼ 3.6 × 10−4 rad s−1.
Our proposal is probably testable with the current Cassini data on gravity
measurements of Titan. The quadrupole Love number has three even az-
imuthal degrees, m = −2, 0, 2, corresponding to three components, k2,−2, k2,0, k2,2.
Equilibrium tides are equally excited at these three azimuthal degrees. How-
ever, g-modes are modified by Titan's rotation and thus cannot be equally
excited at m = −2, 0, 2. Therefore, the three components of the Love number
should be different. We recommend fixing two components of the Love num-
ber at the theoretical k2,eq and fitting the third component. If this algorithm
improves the fitting significantly, it would be a great support for our proposal.
The theoretical k2,eq is a function of the thickness of the outer ice shell (Fig-
9
ure 1). Therefore this procedure may also provide a constraint on d. If our
proposal passes this test, then it is worth carrying out further tests which we
will discuss later.
This introduction section already presents our major results. Readers inter-
ested in more detailed calculations for the amplitude of g-modes and the
Brunt-Vaisala frequency are directed to the next section. Observers interested
in testing our proposal are directed to the final section. Readers interested in
tidal heating mechanisms are directed to Appendix D. We discuss scenarios
involving deep ocean standing waves (g-modes), shallow ocean waves and in-
ertial wave attractors. None of them could produce enough tidal heating to
keep the outer ice shell much thinner than ∼ 100 km in order to enhance the
theoretical k2,eq. However, they apply broadly to satellites of similar size to
Titan which contain internal oceans.
2 Heuristic Treatment of Gravity Modes in Titan's Ocean
Goodman and Dickson (1998) studied gravity waves tidally excited in late-
type stars. The physical situation is similar to Titan. The outer ice shell of
Titan is analogous to the outer convective zone in the star. The presumably
stably-stratified ocean is akin to the inner radiative zone in the star. According
to Goodman and Dickson (1998), at the outer boundary of the stably-stratified
region, the wave's horizontal displacement is similar to that due to equilibrium
tides,
ξwave⊥ ∼ ξeq⊥ , (at r = R − d) ,
(1)
10
if the wave does not reflect back to the outer boundary of the stably-stratified
region. For an incompressible ocean of depth H, the equilibrium tides satisfy
r /H ∼ k⊥ξeq⊥ , and therefore we have
ξeq
ξeq⊥ ∼ ξeq
Hk⊥
r
∼ ξeq
r
R√
ΛH
.
(2)
We assume that the bottom of Titan's ocean is smooth enough so that the
wave reflects at the inner boundary and travels back to where it was excited.
The oscillatory velocity associated with the wave is still in phase with the tidal
force, as long as the travel time, ttra, is shorter than the coherent time, tcoh. In
this case, the amplitude of the horizontal displacment grows by an additional
ξwave⊥ .
The travel time is the time for the wave package to travel through the ocean
radially,
ttra ∼
(cid:90)
ocean
(cid:90)
ocean
dr
vg
∼ 1
ω
drkr ∼ nmode
ω
∼ 1
∆ω
,
(3)
where the radial order, nmode, is the number of half-wavelengths along the
radial direction (Unno et al., 1989),
(cid:90)
nmode ∼ 1
π
(cid:90)
drkr ∼ 1
π
drk⊥
∼
N
ω
ocean
ocean
√
ΛH
πR
N
ω
.
(4)
The frequency separation of g-modes with neighboring radial orders is ∆ω ≡
ωnmode − ωnmode+1 ∼ ω/nmode. We adopt the interior model in Rappaport et al.
(2008). From inside out, Titan contains an inner rocky core, a high-pressure
ice layer, an ocean layer, and an outer ice shell. All parameters are listed in
Table 1. The bottom of the ocean is at Rhp ≈ 2275 km. Given Titan's radius
R ≈ 2575 km and d ∼ 100 km, we have H = R − d − Rhp ∼ 200 km.
11
With N (cid:29) norb ∼ ω, nmode is much greater than unity and thus ∆ω ∼ ω/nmode
is much smaller than ω, which means the spectrum of g-modes is dense near the
tidal frequency. Therefore, the chance is high to find a g-mode (corresponding
to an integer nmode = α) with eigen-frequency ωα ≈ norb . The frequency
mismatch δω ≡ ωα − norb typically falls in the range −0.5∆ω < δω < 0.5∆ω.
In order for the g-mode to enhance k2, the radial displacement of the g-mode
must be in phase with the tidal potential. This requires δω > 0. The chance
for a positive δω is 50% if norb locates randomly in between two neighboring
eigen-frequencies. We do not see a physical reason why a positive δω should
be chosen by Titan. In the case of δω < 0, the g-mode should reduce k2. We
admit this is a caveat for our proposal. Since the current time is probably not
a special one in the evolution history of Titan, norb is not particularly close to
any mode's eigen-frequency. Therefore, we choose δω ∼ 0.5∆ω.
The frequency mismatch would 'drag' the mode out of phase with the tidal
force after a time ∼ π/δω. Damping or scattering of the wave also causes a
phase difference of order π after time ∼ π/γ, where γ is the linear damping
or scattering rate. Nonlinear damping is probably negligible because the tides
are well within the linear regime, ξ⊥k⊥ (cid:28) 1. Therefore, the coherent time,
tcoh, defined as the time it takes for the phase difference to reach π is
tcoh ∼
π
(δω2 + γ2)1/2 ∼ π
δω
,
(5)
where we have assumed γ (cid:28) δω. As long as the boundaries of the ocean layer
are reasonably smooth, this is a good assumption.
Within tcoh, the horizontal displacement of the resonant g-mode increases by
ξwave⊥
every travel time, ttra. After tcoh, the horizontal displacement reaches the
12
saturation value,
ξmode⊥ ∼ ξwave⊥
tcoh
ttra
.
(6)
After a few multiples of π/γ, the g-mode's amplitude is stabilized at this
saturation level. One could understand the statement above using a damped
harmonic oscillator driven near resonance with δω (cid:29) γ.
Combining Equations (1), (2) and (6), we obtain
ξmode⊥ ∼ πR√
ΛH
∆ω
δω
r surf .
ξeq
(7)
In the equation above, we should use ξeq
r at r = R − d. However, with d (cid:28) R,
ξeq
r hardly changes across the ice shell. Thus we adopt the radial displacement
r surf, which is about
of Titan's surface due to equilibrium tides, denoted by ξeq
460 cm.
Gravity modes are almost incompressible, satisfying k⊥ξ⊥ ∼ krξr and k⊥/kr ∼
ω/N (Unno et al., 1989). According to Equation (4), ω/N ∼ √
ΛH
πR ∆ω/ω. Thus
the radial displacement of the g-mode is
ξmode
r
∼ ∆ω2
ω δω
r surf .
ξeq
(8)
The g-mode forces the outer ice shell to move radially by almost the same
amount as demonstrated in Appendix C. The ice shell's rigidity is so weak
compared to Titan's gravity that the shell bends as much as the underlying
ocean 'wants' it to.
In order to enhance the theoretical Love number by (k2,obs−k2,eq)/k2,eq ∼ 46%,
r surf. Thus ξmode
we need the total surface radial displacement to be 1.46 ξeq
r
13
r surf and δω ∼ 0.5∆ω,
needs to be 0.46 ξeq
Equation (8) yields ∆ω ∼ 0.23ω. Combining nmode ∼ ω/∆ω and Equation (4),
r surf. Adopting ξmode
∼ 0.46 ξeq
r
we obtain
Nreq ∼ norb
ω
∆ω
πR√
ΛH
∼ 3.3 × 10−4 rad s−1 ,
(9)
where 'req' means this is the Brunt-Vaisala frequency 'required' to increase
the theoretical k2 by the right amount.
Next, we speculate what physical reason might cause stable stratification in
Titan's ocean. Lunine and Stevenson (1987) proposed a volatile-rich model for
Titan. Their model can refill Titan's atmosphere with methane which other-
wise should have been depleted by photo-chemical reactions. They proposed
that Titan formed by accreting rock and ice in the circum-Saturn nebula. In
the nebula, all the H2O may be combined with volatile molecules as either
Ammonia Monohydrate NH3.H2O or Clathrate CH4.6H2O. The volatile con-
tent of Titan evolves with time (Tobie et al., 2006). Since Titan's atmosphere
is rich in methane now, we speculate the outer ice shell is still composed of
Clathrate. The ice in the deep interior may still remain as Ammonia Monohy-
drate. NH3.H2O is denser than CH4.6H2O by ∆ρ ∼ 0.02 g cm−3 (Stevenson,
1992). There might be a stabilizing composition gradient in the ocean, leading
to
− g
ρ
1
2 ∼
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)comp
(cid:32)
− g
ρ
dρ
dr
∆ρ
H
N ∼
(cid:33) 1
2 ∼ 3.6 × 10−4 rad s−1 ,
(10)
where we employ g ≈ 135 cm s−2 and ρ ≈ 1 g cm−3. This Brunt-Vaisala fre-
quency is similar to Nreq, which may be a coincidence, but at least Nreq is not
an excessively unreasonable value. Questions like whether such a composition
14
gradient is chemically favorable in the ocean, whether it may be destroyed by
convection, and whether there is a salinity gradient or a temperature gradient
as well ask for further investigations. Some of them are difficult to address
because ammonia hydrates under high pressure are poorly understood due to
the lack of experimental data (e.g. Grasset and Pargamin, 2005).
3 Discussion
We discuss predictions testable by observations in this section. All the cal-
culations above neglect Titan's rotation. In reality, Titan is synchronously
rotating, ΩT = norb. Therefore, the quadrupole tidal potential can be sepa-
rated into three components with azimuthal degrees m = −2, 0, 2 respectively.
They have negative, zero and positive azimuthal phase speeds in the rest frame
of Titan. We call them retrograde, axisymmetric and prograde components.
The tidal responses to them can be described by the retrograde, axisymmetric
and prograde Love numbers, denoted by k2,−2, k2,0, k2,2. They are defined in
Appendix B.
Coriolis forces induced by rotation modify the angular patterns of retrograde,
axisymmetric and prograde g-modes differently (Townsend, 2003). The differ-
ence is significant, because ΩT ∼ ω. Therefore, retrograde, axisymmetric and
prograde g-modes are not equally excited. Consequently, their enhancements
to the corresponding components of k2 are different. On the other hand, equi-
librium tides contribute to k2,−2, k2,0 and k2,2 equally, because ΩT (cid:28) ωdyn.
For example, an axisymmetric mode may be preferentially excited, while its
prograde and retrograde counterparts are not. As a result, k2,0 is enhanced by
a mode, whereas k2,2 and k2,−2 are still only due to equilibrium tides. In this
15
case, we expect k2,0 ∼ 0.6 and k2,2 ∼ k2,−2 ∼ k2,eq ∼ 0.42. We examine the
fitting formulas adopted by Iess et al. (2012) and Durante et al. (2019) and re-
alize that they assume k2,0 = k2,2 = k2,−2 = k2. It is true for equilibrium tides
and thus is a natural assumption. Applying fitting formulas distinguishing
the three components of the dynamic Love number would test our proposal.
We recommend adopting k2,eq ≈ 0.42 for two components of k2 and fitting
the third one using the gravity data. If this algorithm improves the fitting
significantly, our proposal would look promising. It is ideal to fit the three
components simultaneously. The largest component may be enhanced by a
g-mode, while the two smaller ones may be the same and thus are due to
equilibrium tides only. Then the latter could be used to indicate the ice shell
thickness using the upper panel of Figure 1.
Moreover, the angular pattern of g-modes along the latitude in a rotating frame
is described by Hough functions instead of associated Legendre polynomials
(Townsend, 2003). Towards large value of ν ≡ 2ΩT/ω, the Hough functions
for g-modes remain close to zero at high latitude with cos θ ≥ ν−1, where θ
is the colatitude. 5 For Titan, ν = 2, the Hough functions concentrate within
latitudes lower than 30◦. The 110th flyby of Titan, labeled by T110 in Durante
et al. (2019), occurred at a high latitude, 74.8◦N. There are three flybys, T022,
T045 and T068, occurring at latitudes greater than 30◦, respectively 45.4◦N,
43.5◦S and 48.9◦S. These flybys are particularly valuable in distinguishing
Hough functions from associated Legendre polynomials.
Because Hough functions are not orthogonal associated Legendre polynomi-
5 There is another set of Hough functions concentrating near the poles. They cor-
respond to r-modes, whose eigen-frequencies are not resonant with norb (Fig.1 in
Townsend, 2003). Therefore we do not discuss them.
16
als, a Hough function with m = 2 usually has non-negligible projections
onto associated Legendre polynomials with (cid:96) greater than two. Therefore, if
prograde or retrograde g-modes are excited, they should lead to time-varying
gravitational potential with angular degrees higher than quadrupole. Because
the gravitational potential decays with the (cid:96) + 1 power of distance from Ti-
tan, flybys at low altitudes are most sensitive to time-varying gravitational
potentials with (cid:96) > 2.
A Equilibrium Tides
We employ the interior model for Titan with an internal ocean presented in
Rappaport et al. (2008). In this model, Titan contains a rocky core, a high-
pressure ice layer , a water ocean layer, and an outermost ice shell, which are
labeled by 'c', 'hp', 'o' and 'i1' respectively. In order to simplify the model, we
assume each layer to be incompressible. We adopt all the parameters on the
left column of Table 7 in Rappaport et al. (2008) except for the outer radius
for the ocean layer, Ro and Bulk modulus. We do not fix Ro. Instead, we set
it to be R − d, where d is the thickness of the outer ice shell which we allow
to vary. Bulk modulus is infinite for an incompressible layer. We then use
this model to solve for the equilibrium tides through the equilibrium equation
(Landau and Lifshit's, 1959),
∇ · σ + ρg = 0 ,
(1)
where σ is the stress tensor and g is the local gravity acceleration. We apply
appropriate boundary conditions, i.e. the Lagrangian perturbation to the rj
(j = r , θ , ϕ) component of the stress tensor must be continuous at each in-
17
terface. The dynamic quadrupole Love number according to equilibrium tides,
k2,eq, of our model is shown by the solid magenta line in the upper panel of
Figure 1. It matches that in Figure 1 of Rappaport et al. (2008) very well.
Note that equilibrium tides have the three components of the quadrupole Love
number the same to each other, and thus we denote them by a single number,
k2,eq.
Equilibrium tides yielded by Φe induces elastic energy, Eeq
elas, which is the
volumetric integration of the inner product of the stress and strain tensors
in the outer ice shell, the high-pressure ice layer and the rocky core. The
total elastic energy. If the peak elastic energy, Eeq
strain and stress tensors vary periodically with frequency norb, and so does the
elaspeak, damps by a fraction,
heat =
1/Q, with frequency norb, the corresponding tidal heating rate is
Eeq
. Note that the definition of Q in this paper refers to peak
(cid:16)
norb
Q
elaspeak
Eeq
(cid:17)
energy instead of time-averaged energy. Because Q must not be smaller than
unity, the optimal
Eeq
heat
with Q = 1 as a function of d is shown by the
(cid:12)(cid:12)(cid:12)max
solid red curve in the lower panel of Figure 1. Also shown in the lower panel
is the conductive heat loss rate Eloss through the outer ice shell of thickness d.
For equilibrium tides, there is a tension between matching k2,eq with k2,obs
and balancing Eloss by Eeq
heat. The former needs a thin shell, although k2,eq is
still more than 1σ below k2,obs at zero shell thickness. The latter requires a
thick shell because the conductive heat loss rate Eloss becomes smaller for a
thicker shell and also a thicker shell means more volume to store elastic energy
which dissipates into heat. Unfortunately for Titan, this tension could not be
resolved by simply adjusting the shell thickness.
This tension is quantitatively illustrated by Figure 1. Equating the optimal
18
(cid:12)(cid:12)(cid:12)max
with Eloss leads to d ∼ 42 km. The corresponding k2,eq ∼ 0.45, which
Eeq
heat
is 2σ below k2,obs. A reasonable Q ≥ 10 requires d ≥ 165 km, making k2,eq ≤
0.385, more than 3σ below the observed Love number. One may artificially
tune the shear modulus for the rocky core and the high-pressure ice layer to
much smaller values. This would enlarge k2,eq as well as Eeq
heat. One does not
want to tune the shear modulus of the outer ice shell to a smaller value because
that would enhance k2,eq only by a tiny amount but make Eeq
heat smaller. We
make µhp and µc ten times smaller and see that this tuning makes the problem
less serious but does not really resolve it, as demonstrated by dashed curves in
Figure 1. We must admit that we do not explore artificial tuning exhaustively,
and therefore we cannot claim confidently that there does not exist a set of
artificially-tuned parameters that makes equilibrium tides work.
B Dynamic Quadrupole Love Number
The time-varying quadrupole tidal potential exerted on Titan by Saturn, to
the leading order of e, is (Murray and Dermott, 1999)
Φe(r, θ, ϕ, t) =
orbr2P 0
3
en2
2
−7
8
1
8
+
2 (cos θ) cos(norbt)
2 (cos θ) cos(2ϕ − norbt)
en2
orbr2P 2
en2
orbr2P 2
2 (cos θ) cos(2ϕ + norbt) .
(2)
(3)
(4)
Here θ and ϕ are the colatitude and longitude of Titan measured in the frame
rotating at Titan's spin rate, ΩT. Titan is synchronized with its orbit, and
thus ΩT = norb. The first term (Equation 2), denoted by Φ2,0
e
, is axisymmetric
because it is independent of ϕ. The second term (Equation 3), denoted by
Φ2,2
e
, is prograde since its azimuthal phase speed is positive. The last term
19
(Equation 4), denoted by Φ2,−2
e
, is retrograde.
The time-varying quadrupole potential induced by the tidal deformation of
Titan in response to Φe may be parameterized as
(r ≥ R, θ, ϕ, t) =
Φself
e
k2,0Φ2,0
e + k2,2Φ2,2
e + k2,−2Φ2,−2
e
(cid:18) R
(cid:19)3(cid:104)
r
(cid:105)
,
r=R
(5)
where k2,0, k2,2 and k2,−2 are the axisymmetric, prograde and retrograde com-
ponents of the dynamic quadrupole Love number.
C Relation between the radial displacements at R and at Ro
We assume the outermost ice shell above the internal ocean is incompressible.
An ocean wave with frequency norb deforms the shell from below. Since the
sound travel time through the ice shell is much shorter than the period of
the ocean wave. The deformation of the ice shell reacts instantaneously to
the ocean wave. Therefore, the equilibrium equation (1) applies. The outer
boundary of the ice shell at r = R is a free surface, and thus the Lagrangian
perturbation to the stress tensor must vanish,
δσ = 0 .
(6)
The lower boundary of the ice shell is in touch with a shear-free ocean, and
thus at r = Ro,
δσrj = 0 , (j = θ, ϕ) .
(7)
The radial displacement must be continuous at r = Ro. Applying these bound-
ary conditions and solving the equilibrium equation, we get a relation between
20
the radial displacement at the surface and that at r = Ro. If the angular shape
of the radial displacement is P 0
2 (cos θ), then the relation is approximately
(cid:33)
(cid:32) ξsurf
ξrr=Ro
r
(cid:32) 4
(cid:33)
≈ 1 − d
20
+
6gρi1
55µi1
11R
≈ 1 − 0.0018
,
(8)
(cid:32) d
(cid:33)
10 km
to the leading order of d. For a different angular shape, the integers in the
bracket after the first ≈ would change, but it is still true that ξsurf
approximated by ξrr=Ro
half-wavelength has a radial displacement ξrr=Ro
. If there is a g-mode excited resonantly and its first
. This mode would force the
is well
r
same radial displacement for the ice shell.
The horizontal displacement at the outer surface of the ice shell takes the
format of ξsurf⊥ R ∇(P 0
2 (cos θ)). We get
(cid:33)
(cid:32) ξsurf⊥
ξrr=Ro
(cid:32) 56
(cid:33)
≈ 3
11
− Rgρi1
55µi1
− d
20
14gρi1
605µi1
121R
≈ 0.255 − 1.88 × 10−3
(cid:32) d
10 km
(cid:33)
.(9)
D Tidal Heating Scenarios other than Equilibrium Tides
If tidal heating could keep a very thin outer ice shell, e.g. d ∼ 10 km, then the
dynamic Love number due to equilibrium tides reaches its maximum value.
Although kmax
2,eq is still smaller than k2,obs, at least the tension between k2,eq
and k2,obs could be partly reconciled. This is also an important question for
understanding Titan's orbital dynamics. Because tidal heat originates from
the epicyclic energy of Titan, tidal heating leads to damping of the orbital
eccentricity.
We have shown in Appendix A that tidal heating rate due to dissipation of
elastic energy in the solid parts of Titan is too small. We propose that g-
21
modes in the presumably stably-stratified ocean may be resonantly excited by
tides. Then it is worth checking whether damping of ocean waves can lead to
any significant tidal heating. The answer is no. We briefly list our order-of-
magnitude estimates below for interested readers.
D.1 Heating due to damping a resonant g-mode
Because g-mode mainly moves horizontally, ξ⊥/ξr ∼ N/norb (cid:29) 1, a g-mode
could potentially contain a lot of kinetic energy. This feature distinguishes
g-modes from equilibrium tides. For the latter, elastic energy greatly exceeds
kinetic energy. However, stable stratification slows down the group velocity
of ocean wave, and thus it takes a long time for a g-mode to grow. In order
to store a lot of kinetic energy in a g-mode, one must give it enough time to
grow. During this time, one cannot damp the mode significantly, otherwise
the mode's growth should have stopped. Although a g-mode can contain a lot
of kinetic energy, this advantage is traded off by its slow damping. Therefore,
tidal heating due to damping of a g-mode is small.
The kinetic energy of a resonant g-mode is roughly
Emode ∼ 4πR2ρoH(norb ξmode⊥
)2 ,
(10)
where ρo is the density of water. Adopting Equations (1), (2), (3), (5) and (6),
we get
ξmode⊥ ∼ ξeq
r surf
πR√
ΛH
∆ω√
δω2 + γ2
.
(11)
The tidal heating rate due to dissipation of Emode is maximized when δω = γ
22
because δω2 + γ2 ≥ 2δωγ. We obtain
mode ∼ (γEmode)max ∼ 2π3(norb ξeq
(cid:21)−1 ∆ω2
Emax
(cid:20) H
∼ 3.0 × 1016
r surf)2norb
erg s−1 ,
R4ρo
ΛH
∆ω2
ωγ
(12)
300 km
r surf ∼ 460 cm which is the maximum radial dis-
placement due to equilibrium tides when d = 0 km. Meanwhile, the mode's
where we have applied ξeq
ωγ
(13)
radial displacement is
ξmode
r
∼ ξeq
r surf
∆ω2√
2γω
.
(14)
Note that the maximal tidal heating rate and the mode's radial displacement
have a common factor ∆ω2/(γω) in their expressions. The observed dynamic
Love number constrains the total surface radial displacement of Titan. There-
2,eq ≈ 1.3. It follows that
fore the factor ξmode
/ ξeq
r
∆ω2/(γω) < 1.8. Thus we get
mode < 5.4 × 1018
Emax
r surf cannot exceed k2,obs/kmax
(cid:20) H
(cid:21)−1
erg s−1 ,
300 km
(15)
which is still one order of magnitude smaller than Eloss ∼ 3.0×1019(d/10 km)−1 erg s−1.
One might say that if the depth of the ocean is 30 km, then the optimal tidal
heating rate due to a g-mode could be big enough to match Eloss. But in order
to fulfill this optimal tidal heating, one has to come up with a physical reason
for δω = γ, i.e. the tidal frequency norb falls particularly close to a g-mode.
Normally we expect the frequency mismatch δω is similar to frequency sepa-
ration of neighboring g-modes, ∆ω, which is much larger than the linewidth
of a g-mode, γ. We find the optimal case implausible although it cannot be
ruled out.
23
D.2 Heating due to resonance ocean wave in a very thin ocean
As the ocean becomes very thin, the dispersion relation for ocean waves be-
comes
ω2 = gk2⊥H .
(16)
Ocean wave is a thin ocean is evanescent radially. It travels horizontally at
the group velocity
∼(cid:113)
vg ∼ ∂ω
∂k⊥
gH .
(17)
The kinetic energy in the ocean wave is
Ewave ∼ 4πR2ρoH(norbξwave⊥ )2 ∼ 4π(norb ξeq
r surf)2 R4ρo
ΛH
,
(18)
wich we have used Equations (1) and (2). The tidal heating rate is maximized
if the ocean wave damps after it travels to the other side of Titan in time
∼ πR/vg ∼ πR/
√
gH. Therefore, the maximal heating rate is
√
wave ∼ Ewave
Emax
gH
πR/
∼ 4R3ρo
Λ
H
(norb ξeq
r surf)2 .
(19)
(cid:18) g
(cid:19)1/2
In order for Emax
wave to balance Eloss ∼ 3 × 1019(d/10 km)−1 erg s−1, one requires
(cid:32) d
(cid:33)2
H ∼ 3.8 × 10−3
10 km
km .
(20)
Such a thin ocean seems very unlikely.
24
D.3
Inertial Wave
Since Titan is rotating, there exists inertial waves (IW) in the ocean. The
restoring force for inertial waves is Coriolis force. According to Ogilvie (2013),
with α = Rc/R ≈ 0.7, we read from their Figure 4. that the imaginary Love
number is at most 2.4 × 10−5 at ω/ΩT = ±1 with the Y 0
2 component. The
corresponding eccentricity damping timescale (Peale, 1999)
(cid:18) a
(cid:19)5
1
1
norb
,
τe =
2
21
MT
MS
R
Im(k2)
(21)
(22)
leads to a tidal heating rate
EIW ∼ MT(anorbe)2
τe
∼ 7 × 1015 erg s−1 .
This is too small to maintain the outer ice shell at a thickness much smallter
than 100 km.
Acknowledgements
We are grateful to Daniele Durante and Luciano Iess who generously shared
their latest observational results. We thank Peter Goldreich, Jeremey Good-
man and Dong Lai for their insightful comments and suggestions. Jing Luan
is supported at the Institute for Advanced Study.
References
R. M. Baland, T. van Hoolst, M. Yseboodt, and O. Karatekin. Titan's obliq-
uity as evidence of a subsurface ocean? Astronomy & Astrophysics, 530:
25
A141, Jun 2011. doi: 10.1051/0004-6361/201116578.
Rose-Marie Baland, Gabriel Tobie, Axel Lef`evre, and Tim Van Hoolst. Titan's
internal structure inferred from its gravity field, shape, and rotation state.
Icarus, 237:29 -- 41, Jul 2014. doi: 10.1016/j.icarus.2014.04.007.
Bruce G. Bills and Francis Nimmo. Rotational dynamics and internal structure
of Titan. Icarus, 214:351 -- 355, Jul 2011. doi: 10.1016/j.icarus.2011.04.028.
Daniele Durante, D. J. Hemingway, P. Racioppa, L. Iess, and D. J. Stevenson.
Titan's gravity field and interior structure after Cassini. Icarus, 326:123 --
132, Jul 2019. doi: 10.1016/j.icarus.2019.03.003.
Jeremy Goodman and Eric S. Dickson. Dynamical Tide in Solar-Type
Binaries. The Astrophysical Journal, 507(2):938 -- 944, Nov 1998.
doi:
10.1086/306348.
O. Grasset and J. Pargamin. The ammonia water system at high pressures:
Implications for the methane of Titan. Planetary and Space Science, 53:
371 -- 384, Apr 2005. doi: 10.1016/j.pss.2004.09.062.
Luciano Iess, Nicole J. Rappaport, Robert A. Jacobson, Paolo Racioppa,
David J. Stevenson, Paolo Tortora, John W. Armstrong, and Sami W. As-
mar. Gravity Field, Shape, and Moment of Inertia of Titan. Science, 327:
1367, Mar 2010. doi: 10.1126/science.1182583.
Luciano Iess, Robert A. Jacobson, Marco Ducci, David J. Stevenson,
Jonathan I. Lunine, John W. Armstrong, Sami W. Asmar, Paolo Racioppa,
Nicole J. Rappaport, and Paolo Tortora. The Tides of Titan. Science, 337:
457, Jul 2012. doi: 10.1126/science.1219631.
L. D. Landau and E. M. Lifshit's. Theory of elasticity. Pergamon Press, 1959.
J. I. Lunine and D. J. Stevenson. Clathrate and ammonia hydrates at high
pressure: Application to the origin of methane on Titan. Icarus, 70:61 -- 77,
Apr 1987. doi: 10.1016/0019-1035(87)90075-3.
26
R. Meriggiola and L. Iess. A new rotational model of Titan from Cassini SAR
data. In European Planetary Science Congress 2012, pages EPSC2012 -- 593,
Sep 2012.
C. D. Murray and S. F. Dermott. Solar system dynamics. Cambridge Univer-
sity Press, 1999.
Gordon I. Ogilvie. Tides in rotating barotropic fluid bodies: the contribution
of inertial waves and the role of internal structure. Monthly Notices of the
Royal Astronomical Society, 429(1):613 -- 632, Feb 2013. doi: 10.1093/mnras/
sts362.
S. J. Peale. Origin and Evolution of the Natural Satellites. Annual Review of
Astronomy and Astrophysics, 37:533 -- 602, Jan 1999. doi: 10.1146/annurev.
astro.37.1.533.
Nicole J. Rappaport, Luciano Iess, John Wahr, Jonathan I. Lunine, J. W.
Armstrong, Sami W. Asmar, Paolo Tortora, Mauro Di Benedetto, and Paolo
Racioppa. Can Cassini detect a subsurface ocean in Titan from gravity
measurements? Icarus, 194:711 -- 720, Apr 2008. doi: 10.1016/j.icarus.2007.
11.024.
P. K. Seidelmann. Explanatory Supplement to the Astronomical Almanac.
University Science Books, 1992.
D. J. Stevenson.
Interior of Titan.
In B. Kaldeich, editor, Symposium on
Titan, volume 338 of ESA Special Publication, April 1992.
G. Tobie, J. I. Lunine, and C. Sotin. Episodic outgassing as the origin of
atmospheric methane on Titan. Nature, 440:61, Mar 2006. doi: 10.1038/
nature04497.
R. H. D. Townsend. Asymptotic expressions for the angular dependence of low-
frequency pulsation modes in rotating stars. Monthly Notices of the Royal
Astronomical Society, 340:1020 -- 1030, Apr 2003. doi: 10.1046/j.1365-8711.
27
2003.06379.x.
W. Unno, Y. Osaki, H. Ando, H. Saio, and H. Shibahashi. Nonradial oscilla-
tions of stars. Tokyo: University of Tokyo Press, 1989.
28
Parameters and Symbols
MT
1.34 × 1023 kg, mass of Titan
R
2575 km, radius of Titan
d
I
e
thickness of the outermost ice shell
0.35◦, orbit inclination with respect to
the equatorial plane of Saturn
orbital eccentricity, current value enow ≈ 0.0288
norb
4.56 × 10−6 rad s−1, orbit mean motion
a
1.22 × 106 km, orbit semi-major axis
ρi1
ρo
ρhp
ρave
0.92 g cm−3
1.0 g cm−3
1.31 g cm−3
1.881 g cm−3
µi1
3.3 GPa
µo
0.0 GPa
µhp
4.6 GPa
µc
60.0 GPa
k2,obs
0.616 ± 0.067, measured quadrupole Love number of Titan Rhp
2275 km
kice
cp
2 × 105 erg K−1 cm−1 s−1, thermal conductivity of ice
2 kJ kg−1 K−1, specific heat capacity of ice
Rc
1670 km
Ro R − d
∆T
180 K, temperature contrast through the ice shell
g
135 cm s−2, surface gravity of Titan
Table 1
The right panel quotes Table 7 of Rappaport et al. (2008). The subscripts 'i1', 'o',
'hp' and 'c' label the outer most ice shell, the liquid ocean layer, the high-pressure
ice layer and the rocky core respectively. Ro refers to the outer radius of the ocean
layer, etc. ρ denotes density and µ shear modulus. As the thickness of the outer ice
shell is varied, the density of the core is adjusted to keep the same average density,
ρave.
29
Fig. 1. This figure shows that equilibrium tides of Titan could not account for the
observed k2 and generate enough tidal heating to maintain a thin outer ice shell at
the same time. The upper panel shows the theoretical k2 due to equilibrium tides of
Titan as a function of the thickness of the outermost ice shell, d. The solid magenta
line adopts the normal values for shear modulus as shown in Table 1. The dashed
magenta line denoted by 'soft core' artificially scales down the shear modulus of the
high-pressure ice layer and the rocky core by a factor of 10. The solid black horizontal
line labels the measured k2,obs = 0.616, and the grey horizontal bar labels the 1σ
uncertainty, δk2,obs = 0.067. The lower panel shows the maximum tidal heating rate
produced by dissipation of elastic energy due to equilibrium tides adopting a tidal
quality factor, Q = 1, by the solid red line. The denotation 'soft core' has the same
meaning as that in the upper panel. The blue curve shows the conductive heat loss
rate,
Eloss ∼ 4πR2kice∆T /d ∼ 3.0 × 1019(d/10 km)−1 erg s−1.
30
204060801000.400.400.450.450.500.500.550.550.600.600.650.650.700.70k2 soft coreobserved k220406080100ice shell thickness d (km)101101100100Heat Generation v.s. Loss Rates 1019ergs1soft core, Q=1Q=1 |
1911.11953 | 1 | 1911 | 2019-11-27T04:55:38 | OGLE-2016-BLG-1227L: A Wide-separation Planet from a Very Short-timescale Microlensing Event | [
"astro-ph.EP",
"astro-ph.SR"
] | We present the analysis of the microlensing event OGLE-2016-BLG-1227. The light curve of this short-duration event appears to be a single-lens event affected by severe finite-source effects. Analysis of the light curve based on single-lens single-source (1L1S) modeling yields very small values of the event timescale, $t_{\rm E}\sim 3.5$ days, and the angular Einstein radius, $\theta_{\rm E}\sim 0.009$ mas, making the lens a candidate of a free-floating planet. Close inspection reveals that the 1L1S solution leaves small residuals with amplitude $\Delta I\lesssim 0.03$ mag. We find that the residuals are explained by the existence of an additional widely-separated heavier lens component, indicating that the lens is a wide-separation planetary system rather than a free-floating planet. From Bayesian analysis, it is estimated that the planet has a mass of $M_{\rm p} = 0.79^{+1.30}_{-0.39} M_{\rm J}$ and it is orbiting a low-mass host star with a mass of $M_{\rm host}=0.10^{+0.17}_{-0.05} M_\odot$ located with a projected separation of $a_\perp=3.4^{+2.1}_{-1.0}$ au. The planetary system is located in the Galactic bulge with a line-of-sight separation from the source star of $D_{\rm LS}=1.21^{+0.96}_{-0.63}$ kpc. The event shows that there are a range of deviations in the signatures of host stars for apparently isolated planetary lensing events and that it is possible to identify a host even when a deviation is subtle. | astro-ph.EP | astro-ph |
DRAFT VERSION NOVEMBER 28, 2019
Preprint typeset using LATEX style emulateapj v. 12/16/11
OGLE-2016-BLG-1227L: A WIDE-SEPARATION PLANET FROM A VERY SHORT-TIMESCALE MICROLENSING
EVENT
CHEONGHO HAN0001, ANDRZEJ UDALSKI0003,100, ANDREW GOULD0004,0005,101
(LEADING AUTHORS),
MICHAEL D. ALBROW0007, SUN-JU CHUNG0002,0008, KYU-HA HWANG0002, YOUN KIL JUNG0002, CHUNG-UK LEE0002,101,
YOON-HYUN RYU0002, IN-GU SHIN0002, YOSSI SHVARTZVALD0009, JENNIFER C. YEE0010, WEICHENG ZANG0011,
SANG-MOK CHA0002,0012, DONG-JIN KIM0002, HYOUN-WOO KIM0002, SEUNG-LEE KIM0002,0008, DONG-JOO LEE0002,
YONGSEOK LEE0002,0012, BYEONG-GON PARK0002,0008, RICHARD W. POGGE0005, M. JAMES JEE0013,0014, DOEON KIM0001,
AND
CHUN-HWEY KIM0016, WOONG-TAE KIM0017
(THE KMTNET COLLABORATION),
PRZEMEK MRÓZ0003,0018, MICHAŁ K. SZYMA ´NSKI0003, JAN SKOWRON0003, RADEK POLESKI0005, IGOR SOSZY ´NSKI0003,
PAWEŁ PIETRUKOWICZ0003, SZYMON KOZŁOWSKI0003, KRZYSZTOF ULACZYK0015
(THE OGLE COLLABORATION)
0001 Department of Physics, Chungbuk National University, Cheongju 28644, Republic of Korea; [email protected]
0002 Korea Astronomy and Space Science Institute, Daejon 34055, Republic of Korea
0003 Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
0004 Max Planck Institute for Astronomy, Königstuhl 17, D-69117 Heidelberg, Germany
0005 Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA
0007 University of Canterbury, Department of Physics and Astronomy, Private Bag 4800, Christchurch 8020, New Zealand
0008 Korea University of Science and Technology, 217 Gajeong-ro, Yuseong-gu, Daejeon, 34113, Republic of Korea
0009 Department of Particle Physics and Astrophysics, Weizmann Institute of Science, Rehovot 76100, Israel
0010 Center for Astrophysics Harvard & Smithsonian 60 Garden St., Cambridge, MA 02138, USA
0011 Department of Astronomy and Tsinghua Centre for Astrophysics, Tsinghua University, Beijing 100084, China
0012 School of Space Research, Kyung Hee University, Yongin, Kyeonggi 17104, Republic of Korea
0013 Yonsei University, Department of Astronomy, Seoul, Republic of Korea
0014 Department of Physics, University of California, Davis, California, USA
0015 Department of Physics, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK
0016 Department of Astronomy & Space Science, Chungbuk National University, Cheongju 28644, Republic of Korea
0017 Department of Physics & Astronomy, Seoul National University, Seoul 08826, Republic of Korea and
0018 Division of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena, CA 91125, USA
Draft version November 28, 2019
ABSTRACT
We present the analysis of the microlensing event OGLE-2016-BLG-1227. The light curve of this short-
duration event appears to be a single-lens event affected by severe finite-source effects. Analysis of the light
curve based on single-lens single-source (1L1S) modeling yields very small values of the event timescale, tE ∼
3.5 days, and the angular Einstein radius, θE ∼ 0.009 mas, making the lens a candidate of a free-floating planet.
Close inspection reveals that the 1L1S solution leaves small residuals with amplitude ∆I . 0.03 mag. We find
that the residuals are explained by the existence of an additional widely-separated heavier lens component,
indicating that the lens is a wide-separation planetary system rather than a free-floating planet. From Bayesian
analysis, it is estimated that the planet has a mass of Mp = 0.79+1.30
−0.39 MJ and it is orbiting a low-mass host star
with a mass of Mhost = 0.10+0.17
−1.0 au. The planetary system
is located in the Galactic bulge with a line-of-sight separation from the source star of DLS = 1.21+0.96
−0.63 kpc. The
event shows that there are a range of deviations in the signatures of host stars for apparently isolated planetary
lensing events and that it is possible to identify a host even when a deviation is subtle.
Subject headings: gravitational lensing: micro -- planetary systems
−0.05 M⊙ located with a projected separation of a⊥ = 3.4+2.1
1. INTRODUCTION
Although most microlensing planets are detected through
the channel of a short-term perturbation to the standard lens-
ing light curve of the planet host (Mao & Paczy´nski 1991;
Gould & Loeb 1992), a fraction of planets can be detected
through the channel of an isolated lensing event produced
by the gravity of the planet itself (Bennett & Rhie 2002;
Han et al. 2004). The latter channel is important because
it provides a unique method to probe free-floating planets
(FFPs) that may have been ejected from the planetary systems
in which they formed or have not been gravitationally bound
to any host star before.
The most important characteristics of an FFP lensing event
is its short timescale. This is because the event timescale tE is
related to the angular Einstein radius θE and the relative lens-
source proper motion µ by tE = θE/µ, and the angular Einstein
radius is proportional to the square root of the lens mass M,
i.e.,
100 OGLE Collaboration.
101 KMTNet Collaboration.
θE = (κMπrel)1/2,
πrel = au(cid:18) 1
DL
−
1
DS(cid:19) .
(1)
2
Han et al.
TABLE 1
DATA USED IN ANALYSIS
Data set
OGLE
KMTC
KMTS
Ndata
154
369
575
Range (HJD′)
7110.8 -- 7659.6
7500.7 -- 7599.7
7441.6 -- 7675.3
NOTE. -- Ndata indicates the number of each data set.
caustic induced by the binarity of the planet-host system. For
a binary lens composed of a planet and a host, there exist two
sets of caustics. One set of caustics is located close to the
host (central caustic) and the other caustic (planetary caustic)
is located at a distance of sc = s − 1/s from the host. Here s
represents the projected planet-host separation normalized to
θE. The planetary caustic of a wide-separation planet forms a
closed curve with 4 cusps. The full width along the star-planet
axis, ∆ξc, and the height normal to the star-planet axis, ∆ηc,
of the caustic are
∆ξc =
4q1/2
s√s2 − 1
;
∆ηc =
4q1/2
s√s2 + 1
(2)
respectively (Han 2006a). For a wide-separation planet with
s ≫ 1, the planetary caustic is located close to the planet, i.e.,
sc → s, and both ∆ξc and ∆ηc approaches 4q1/2s−2, forming
an astroid-shape caustic. The caustic size rapidly shrinks with
the increase of the planet-host separation, i.e., ∆ξc ∼ ∆ηc ∝
s−2. As the caustic becomes smaller, the signature of the host
star diminishes with the increasing finite-source effects.
In this paper, we present the analysis of the lensing event
OGLE-2016-BLG-1227. The light curve of the event appears
to be approximated by a short-timescale 1L1S model with se-
vere finite-source effects, making the lens a candidate FFP.
From the close inspection of the light curve, it is found that the
1L1S solution leaves small residuals. We inspect the origin
of the residuals to check the existence of a widely-separated
heavier lens component, i.e., host of the planet.
We organize the paper as follows. In Section 2, we describe
the observations of the lensing event and the data obtained
from these observations. In Section 3, we present the analysis
of the event based on the 1L1S interpretation. In Section 4, we
inspect the possible existence of a widely separated host of the
planet by conducting a binary-lens (2L1S) analysis. In Sec-
tion 5, we estimate the angular Einstein radius by determining
the dereddened color and brightness of the source star. In Sec-
tion 6, we conduct Bayesian analysis of the event to determine
the physical lens parameters including the mass and location
of the lens system. We summarize the results and conclude in
Section 7.
2. OBSERVATION AND DATA
The lensing event OGLE-2016-BLG-1227 occurred on a
star located toward the Galactic bulge field. The equatorial
coordinates of the lensed star (source) are (R.A.,decl.)J2000 =
(17 : 42 : 23.31, −33 : 45 : 35.2), which correspond to the
galactic coordinates (l, b) = (−4◦.47, −1◦.94). The source of
the event is a bright giant with a baseline magnitude of Ibase =
16.89 from the calibrated OGLE photometric maps.
is detecting the blended light from a host star by conducting high-resolution
observations (Bennett & Rhie 2002). The last proposed method is conducting
astrometric follow-up observations of isolated events using high-precision in-
terferometers (Han 2006b).
FIG. 1. -- Lightcurve of OGLE-2016-BLG-1227. The curve drawn on the
data points is the model obtained from the 1L1S fitting to the light curve
considering finite-source effects.
Here κ = 4G/(c2au), πrel represents the relative lens-source
parallax, and DL and DS denote the distances to the lens and
source, respectively. For FFP events, the chance to exhibit
deformed lensing light curves caused by severe finite-source
effects is high. The deformation can occur when the angular
source radius θ∗ is comparable to θE for FFP event, in which
case the light curve is very likely to be affected by severe
finite-source effects. Recently, three candidates of FFPs were
reported by Mróz et al. (2018) and Mróz et al. (2019) from
the analyses of the lensing events with these characteristics.
However, even when an event is both very short and ex-
hibits strong finite-source effects, the lens cannot be securely
identified as an "FFP". First, it is always possible that the
small value of θE derives from a small πrel rather than small
lens mass M. See Equation (1). This issue can only be re-
solved for individual FFP candidates by measuring the mi-
crolens parallax πE ≡ πrel/θE, using, e.g., a satellite in solar
orbit (Refsdal 1966) or so-called terrestrial parallax (Gould
1997; Gould et al. 2009). Nevertheless, from an ensemble
of θE measurements (even without corresponding πE mea-
surements), one can statistically constrain the properties of
the FFP population. However, there is a second fundamen-
tal problem that has the potential to corrupt such a statistical
sample, namely that a wide-separation planet can also pro-
duce a lensing light curve with similar characteristics, mas-
querading as an FFP. Therefore, it is important to distinguish
the two populations of events produced by FFPs and wide-
separation planets in order to draw statistically meaningful
conclusions about the properties and frequency of both bound
and unbound planets.
Han & Kang (2003) pointed out that an important frac-
tion of isolated short-timescale events produced by wide-
separation planets can be distinguished from those produced
by FFPs by detecting the signatures of host stars in the lens-
ing light curves.3 The signatures arise due to the planetary
3 Besides this method, the nature of a wide-separation planet can be iden-
tified by several other methods. One method is detecting long-term bumps in
the light curve caused by the primary star (Han et al. 2005). Another method
Wide-separation Microlensing Planet
3
TABLE 2
LENSING PARAMETERS
Parameter
1L1S
Inner solution
2L1S
Outer solution
χ2
t0 (HJD′)
u0
tE (days)
tE,1 (days)
tE,2 (days)
s
q
α (rad)
ρ
teff = u0tE (days)
t∗ = ρtE (days)
tp = q−1/2tE (days)
1115.1
7561.920 ± 0.017
0.681 ± 0.017
3.54 ± 0.05
1.05 ± 0.013
--
--
--
--
--
--
--
--
968.6
7561.999 ± 0.031
0.066 ± 0.012
45.37 ± 8.07
4.05 ± 0.06
45.19 ± 8.09
3.68 ± 0.21
124.48 ± 46.79
4.783 ± 0.062
0.092 ± 0.017
3.00 ± 0.08
4.17 ± 0.03
4.07 ± 0.06
973.0
7561.976 ± 0.032
−0.057 ± 0.012
52.23 ± 12.76
4.01 ± 0.06
52.07 ± 12.79
3.57 ± 0.24
168.99 ± 98.86
4.689 ± 0.066
0.080 ± 0.018
2.97 ± 0.08
4.16 ± 0.03
4.02 ± 0.06
NOTE. -- HJD′ = HJD − 2450000. For the 2L1S solution, tE represents the event timescale corresponding to the total mass of the binary lens, and
tE,1 and tE,2 represent the timescales corresponding to the masses of individual lens components, M1 and M2, respectively. The subscripts of the lens
components are chosen according to the distances from the source trajectory. The source trajectory passes closer to the lower-mass lens component
and thus M1 < M2, tE,1 < tE,2, and q = M2/M1 > 1.
globally distributed in the southern hemisphere at the Sid-
ing Spring Observatory in Australia (KMTA), Cerro Tololo
Interamerican Observatory in Chile (KMTC), and the South
African Astronomical Observatory in South Africa (KMTS).
Each KMTNet telescope is equipped with a camera, consist-
ing of four 9k× 9k chips, yielding 4 deg2 field of view. The
event was found from the analysis of the data conducted af-
ter the 2016 season (Kim et al. 2018) and it was designated as
KMT-2016-BLG-1089. Most KMTNet images were obtained
in I band and about one tenth of images were obtained in V
band for the source color measurement. Thanks to the high-
cadence coverage (1 hr−1 for each telescope) using the multi-
ple telescopes, the detailed structure of the light curve is well
delineated by the KMTNet data, despite the short duration of
the event.
Reduction of the data was carried out using the photometry
codes developed by the individual survey groups: Wo´zniak
(2000) for the OGLE and Albrow et al. (2009) for the KMT-
Net data sets. These codes are based on the difference imag-
ing method developed by Alard & Lupton (1998). For a sub-
set of the KMTNet data sets, additional photometry is con-
ducted using the pyDIA code (Albrow 2017) to measure the
source color. The errorbars of the individual data sets are
readjusted according to the procedure described in Yee et al.
(2012). We note that the KMTA data set is not used in the
analysis because the photometry quality is relatively low and
the data do not cover the major part of the light curve.
In
Table 1, we list the data sets used in the analysis along with
numbers of data points, Ndata, and the time ranges of the indi-
vidual data sets.
3. SINGLE-LENS SINGLE-SOURCE (1L1S) MODELING
In Figure 1, we present the light curve of OGLE-2016-
BLG-1227. The light curve appears to be that of a 1L1S event
affected by severe finite-source effects. We, therefore, start
the analysis of the event by conducting a 1L1S modeling.
The modeling is carried out by searching for the lensing
parameters that best describe the observed light curve. The
light curve of a 1L1S event affected by finite-source effects
is described by four lensing parameters. These parameters
include the time of the closest lens-source approach, t0, the
lens-source separation at that time, u0, the event timescale, tE,
and the normalized source radius, ρ. The normalized source
FIG. 2. -- Comparison of the lensing lightcurve with those of four compari-
son stars around the lensing source. The lower four panels show the residuals
of the comparison stars from baseline magnitudes and the second panel shows
the residuals of the lensing event from the 1L1S solution.
The lensing event was first discovered by the Optical Gravi-
tational Lensing Experiment (OGLE: Udalski et al. 2015) sur-
vey, and the discovery was notified to the microlensing com-
munity on 2016 June 29. The OGLE survey was conducted
utilizing the 1.3 m telescope located at the Las Campanas Ob-
servatory in Chile. The telescope is equipped with a camera,
which consists of 32 2k× 4k chips, yielding a 1.4 deg2 field
of view. The OGLE images were obtained mostly in I band
and some images were taken in V band for the source color
measurement.
The event was also located in the field toward which the
Korea Microlensing Telescope Network survey (KMTNet:
Kim et al. 2016) was monitoring. The KMTNet survey was
conducted using the three identical 1.6 m telescopes that are
4
Han et al.
FIG. 3. -- Comparison of the 1L1S (dotted curve) and 2L1S (solid curve) solutions. The middle and bottom panels show the residuals from the 1L1S and
2L1S solutions, respectively. The solid curve in the middle panel represents the difference between the 1L1S and 2L1S solutions. In the top panel, the arrows at
t1(HJD′) = 7559.3 and t2(HJD′) = 7564.6 represent the times of the two dips in the residuals from the 1L1S solution.
radius is defined as the ratio of the angular source radius θ∗
to the angular Einstein radius, i.e., ρ = θ∗/θE, and it is needed
to describe the deformed light curve caused by finite-source
effects. We search for the best-fit lensing parameters using
the Markov Chain Monte Carlo (MCMC) method.
In computing finite-source magnifications, we consider the
variation of the source surface brightness caused by limb
darkening (Witt 1995; Valls-Gabaud 1995; Loeb & Sasselov
1995). To account for the limb-darkening variation, we model
the surface brightness of the source star as
Sλ = ¯Sλ(cid:20)1 − Γλ(cid:18)1 −
3
2
cos φ(cid:19)(cid:21) ,
(3)
where ¯Sλ denotes the mean surface brightness, Γλ is the lin-
ear limb-darkening coefficient, and φ represents the angle be-
tween the line of sight toward the center of the source star and
the normal to the source surface. The limb-darkening coef-
ficient is determined based on the stellar type of the source
star. As we will show in Section 5, the source is a bulge gi-
ant with a spectral type K3. Based on the stellar type, we
set the limb-darkening coefficient as ΓI = 0.41 and ΓV = 0.74
by adopting the values from Claret (2000) under the assump-
tion that vturb = 2 km s−1, log(g/g⊙) = −2.4, and Teff = 4500 K.
For the computation of finite-source magnifications, we use
the semianalytic expressions derived by Gould (1994) and
Witt & Mao (1994).
In Table 2, we present the best-fit lensing parameters ob-
tained from the 1L1S modeling. In Figure 1, we also present
the model curve superposed on the data points. We note that
the estimated event timescale, tE ∼ 3.5 days, is much shorter
than those of typical lensing events with ∼ (O)10 days al-
though events with such short timescales are not extremely
rare. Furthermore, the normalized source radius, ρ ∼ 1.05, is
much bigger than typical values of ∼ 0.01 -- 0.02 for events
involved with giant source stars. The unusually large ρ value
suggests that the angular Einstein radius is likely to be very
small. As we will show in Section 5, the angular radius of the
source is θ∗ ∼ 9.0 µas, and thus the angular Einstein radius of
the event is θE ∼ 0.009 mas. This is very much smaller than
∼ 0.5 mas of typical lensing events. The very small values of
tE and θE make the lens of the event a candidate of an FFP or a
brown-dwarf. We note that the lens of the event was originally
found as a brown-dwarf or an FFP candidate from the search
for isolated events with short tE and very small θE conducted
by Han et al. (2019), but the analysis is separately presented
in this work for the reason presented in Section 4.
Although the observed light curve appears to be approxi-
mated by the 1L1S model, it is found that the solution leaves
small residuals with amplitude ∆I . 0.03 mag. See the lower
panel of Figure 1. The source was located close to the Moon
during the lensing magnification and thus the photometry We
check this possibility by conducting additional photometry for
nearby stars. In Figure 2, we present the lightcurves of four
comparison stars and compare them with that of the lensing
event. It shows that the magnitudes of the comparison stars
remain constant in contrast to the 1L1S residuals. This indi-
cates that the photometry is not affected by the Moon and the
residuals from the 1L1S solution are likely to be real.
4. BINARY-LENS SINGLE-SOURCE (2L1S) MODELING
Considering that the main part of the lensing light curve is
produced by a planetary-mass object, we check whether there
exists a host star located away from the planet. For this, we
Wide-separation Microlensing Planet
5
FIG. 4. -- Cumulative distribution of ∆χ2 between the 1L1S and 2L1S
models (lower panel). The light curve in the upper panel is presented to show
the region of fit improvement.
additionally conduct a 2L1S modeling of the light curve.
Compared to the 1L1S modeling, the 2L1S modeling re-
quires three additional lensing parameters to describe the lens
binarity. These parameters include the projected binary sep-
aration normalized to the angular Einstein radius, s, the mass
ratio between the lens components, q = M2/M1, and the angle
between the binary axis and the source trajectory, α (source
trajectory angle).
In the 2L1S modeling, the solution of the lensing parame-
ters is searched for in two steps. In the first step, we conduct a
grid search for the parameters s and q, while the other param-
eters are searched for using the MCMC method. This proce-
dure yields a χ2 map on the s -- q parameter plane and we find
local minima that appear in the map. In the second step, we
refine the individual local minima by additionally conducting
modeling with all parameters, including the grid parameters s
and q, allowed to vary. We find a global solution by compar-
ing the goodness of the local solutions. This procedure allows
us to find degenerate solutions, if they exist.
We find that the model fit substantially improves with the
introduction of an additional widely-separated lens compo-
nent M2. The additional lens component has a mass much
heavier than the lens component M1 responsible for the short
magnified part of the light curve, suggesting that the addi-
tional lens component is the host of the planet. In Figure 3,
we present both the 1L1S and 2L1S models and the residu-
als from the individual models. The solid curve superposed
on the residuals of the 1L1S model in the middle panel rep-
resents the difference between the 1L1S and 2L1S models. It
is found that the 2L1S residuals are substantially reduced rel-
ative to the 1L1S model. In Figure 4, we present the cumula-
tive distribution of ∆χ2 = χ2
2L1S between the 1L1S and
2L1S models to better show the region of the fit improvement.
We find that the 2L1S improves the fit by ∆χ2 ∼ 146.5. We
further check whether there is an additional weak long-term
bump caused by the heavier companion, but we find no such
a bump. As we will show below, the reason for the absence of
1L1S − χ2
FIG. 5. -- Lens-system configurations of the inner and outer 2L1S solu-
tions. Coordinates are centered at the center of the planetary caustic. The
time t0 is the time of the closest source approach to the planetary caustic, and
the times t1 and t2 correspond to the times of the two dips in the residuals
from the 1L1S model presented in Fig. 4. In each panel, the line with an
arrow represents the source trajectory. The circles on the source trajectory in
the two lower panels represent the source positions at t0, t1, and t2. The size
of the circle is scaled to the source size.
a bump is that the source passes perpendicular to the binary
axis.
In searching for lensing solutions, we find that the observed
light curve is subject to the so-called "inner/outer degener-
acy". This degeneracy arises because the planetary anomalies
produced by the source approaching the inner and outer sides
(with respect to the host of the planet) of the planetary caustic
are similar to each other (Gaudi & Gould 1997). It is found
that the degeneracy is severe although the inner solution is
slightly preferred over the outer solution by ∆χ2 ∼ 4.3.
In Table 2, we list the best-fit lensing parameters of the
2L1S solutions for both the inner and outer solutions. For
each solution, we present three values of timescales (tE, tE,1,
tE,2), in which tE represents the event timescale corresponding
to the total mass of the binary lens, and tE,1 = [1/(1 + q)]1/2tE
and tE,2 = [q/(1 + q)]1/2tE represent the timescales correspond-
ing to the masses of individual lens components, M1 and M2.
We note that the subscripts of the lens components M1 and
M2 are chosen according to the distances from the source
trajectory. The source trajectory approaches closer to the
lower-mass lens component and thus M1 < M2, tE,1 < tE,2,
and q = M2/M1 > 1. The estimated mass ratio between the
lens components, q ∼ 124 for the inner solution and q ∼ 169
for the outer solution, is much bigger than unity, indicating
that M2 is the host of the planet M1. The host is separated
from the planet with a projected separation of s ∼ 3.6.
In Figure 5, we present the lens-system configurations of
the inner and outer 2L1S solutions. The upper panel shows the
whole view including the both lens components. The lower
two panels show the zoom of the region around the plane-
tary caustic for the inner (right panel) and outer (left panel)
solutions. The three brown-tone circles in the lower panels
represent the source positions at three different times of t0,
t1, and t2. The time t0 corresponds to the time of the clos-
est source approach to the planetary caustic, and the times
t1(HJD′) = 7559.3 and t2(HJD′) = 7564.6 correspond to the
times of the two dips in the residuals from the 1L1S model.
See the corresponding times t1 and t2 marked in Figure 3. The
size of the circles is scaled to the source size. It is found that
6
Han et al.
FIG. 6. -- ∆χ2 distributions of points in the MCMC chain on the parameter
planes of the (teff,t∗,tp) combinations. The red, yellow, green, and blue colors
represent points with 1σ, 2σ, 3σ, and 4σ, respectively. The distributions are
constructed based on the "inner 2L1S solution".
FIG. 7. -- Source location (blue empty circle) with respect to the centroid of
red giant clump (RGC, red dots) in the instrumental color-magnitude diagram
constructed based on the pyDIA photometry of the KMTC data set.
the source is much bigger than the caustic. This causes severe
attenuation of the signal induced by the caustic and makes the
light curve appear to be very similar to that of a 1L1S event.
We note that the estimated lensing parameters have large
uncertainties. See Table 2. The main reason for the large un-
certainties of the lensing parameters is that the observed lens-
ing magnification is mostly produced by the planet, and the
planet's host is characterized by the subtle deviations in the
planet-induced magnifications. In this case, the uncertainty
of the timescale tE ∼ tE,2 is large. The large uncertainty of
tE propagates into the uncertainty of mass ratio because the
mass ratio is related to the timescale by q = (tE,1/tE,2)1/2 ∼
(tE,1/tE)1/2. The uncertain timescale also induces large un-
certainties of u0 and ρ because the measured caustic-crossing
duration results from the combination of these parameters by
tcc = 2(u2
0 + ρ2)1/2tE.
In Figure 6, we present the ∆χ2 distributions of points in
the MCMC chain on the teff -- t∗ -- tp parameter planes. The in-
dividual timescales represent teff = u0tE, t∗ = ρtE, and tp =
q−1/2tE, respectively. The "effective timescale" teff is fre-
quently used because it facilitates intuitive understanding of
a light curve independent of separately determining u0 and
tE from modeling. The "source-crossing timescale" t∗ rep-
resents an approximate timescale for the lens to transit the
source surface. Finally, the "planet timescale" tp denotes an
approximate timescale of the isolated event produced by the
planet. We present the estimated values of these timescales
in Table 2. These timescales are derived from the shape of
a lensing light curve, and thus they are tightly constrained
despite the large uncertainties of the lensing parameters, as
demonstrated in Figure 6.
5. ANGULAR EINSTEIN RADIUS
We determine the angular Einstein radius from the normal-
ized source radius ρ together with the angular source radius
θ∗ by θE = θ∗/ρ. The normalized source radius is deter-
mined from modeling the light curve. For the estimation of
the angular source radius, we use the method of Yoo et al.
(2004). According to this method, we first place the source
position in the instrumental color-magnitude diagram (CMD)
of stars around the source. We then measure the offsets in
color, ∆(V − I), and magnitude, ∆I, of the source from the
centroid of the red giant clump (RGC) in the CMD. With
the measured offsets ∆(V − I) and ∆I together the known
dereddened source color and magnitude of the RGC centroid,
(V − I, I)RGC,0 = (1.06,14.65) (Bensby et al. 2013; Nataf et al.
2013), the dereddened color and magnitude of the source are
estimated by
(V − I, I)0 = (V − I, I)RGC,0 + ∆(V − I, I).
(4)
In Figure 7, we present the positions of the source and
the RGC centroid in the instrumental CMD. The CMD is
constructed using the pyDIA photometry of the KMTC data
set. We note that the location of the blend cannot be deter-
mined because the baseline flux is dominated by the source
flux and the flux from the blend is consistent with zero
within the photometry uncertainty. The color and magni-
tude of the source in the instrumental CMD are (V − I, I) =
(3.91 ± 0.11,16.85 ± 0.01) compared to those of the RGC
centroid of (V − I, I)RGC = (3.59,17.43). With the measured
offsets of ∆(V − I) = 0.32± 0.11 and ∆I = 0.58± 0.01, the de-
reddened color and brightness of the source are estimated as
(V − I, I)0 = (1.38± 0.11,14.07± 0.01). The estimated source
color and brightness indicate that the source is a typical bulge
giant with a spectral type K3.
Once the dereddened color and magnitude are determined,
we then estimated the angular source radius. For this, we first
convert the V − I color into V − K color using the color-color
relation of Bessell & Brett (1988) and then the angular source
radius is estimated using the Kervella et al. (2004) relation be-
tween V − K and θ∗. This procedure yields an angular source
radius of
θ∗ = 9.01± 1.15 µas.
(5)
Wide-separation Microlensing Planet
7
ANGULAR EINSTEIN RADIUS AND RELATIVE LENS-SOURCE PROPER
TABLE 3
Parameter
θE (mas)
θE,1 (mas)
θE,2 (mas)
µ (mas yr−1)
MOTION
Inner solution
0.098 ± 0.044
0.007 ± 0.003
0.097 ± 0.043
0.79 ± 0.10
Outer solution
0.113 ± 0.058
0.009 ± 0.004
0.112 ± 0.057
0.79 ± 0.10
NOTE. -- The Einstein radius θE corresponds to the total mass of the
lens M = M1 + M2, and θE,1 and θE,2 represent the Einstein radii correspond-
ing to M1 and M2, respectively.
With the measured angular source radius, the angular Einstein
radius is estimated as
θE =
θ∗
ρ
=(cid:26)0.098± 0.044 mas
0.113± 0.058 mas
(inner solution),
(outer solution),
The estimated relative lens-source proper motion is
µ =
θE
tE
= 0.79± 0.10 mas yr−1
(6)
(7)
for both the inner and outer solutions. We note that the frac-
tional uncertainty of the relative lens-source proper motion,
σµ/µ ∼ 13%, is substantially smaller than the uncertainty
of the angular Einstein radius, σθE/θE ∼ 50%. This is be-
cause the proper motion in the lensing modeling is computed
by µ ∼ θ∗/t∗ and the uncertainty of the "source-crossing
timescale" t∗ is significantly smaller than the uncertainty of
the event timescale tE.
In Table 3, we summarize the estimated Einstein radii and
relative lens-source proper motions for the inner and outer
solutions. Also presented are the angular Einstein radii cor-
responding to the masses of the individual lens components,
θE,1, and θE,2, similar to the presentation of tE,1 and tE,2 in Ta-
ble 2. We note that the estimated θE,1 ∼ 0.007 -- 0.009 mas is
consistent with the Einstein radius estimated from the 1L1S
modeling. We also note that the measured angular Einstein
radius, θE ∼ 0.1 mas, is substantially smaller than ∼ 0.5 mas
of a typical lensing event produced by a low-mass star with a
mass of ∼ 0.3 M⊙ located roughly halfway between the ob-
server and the bulge source. The angular Einstein radius is
related to the lens mass and distance by Equation (1). Then,
the small angular Einstein radius suggests that the lens has a
small mass and/or it is located close to the source.
6. PHYSICAL LENS PARAMETERS
For the unique determinations of the physical lens param-
eters of the lens mass M and distance DL, one must measure
both the angular Einstein radius and the microlens parallax
πE, i.e.,
M =
;
DL =
.
(8)
θE
κπE
au
πEθE + πS
Here πS = au/DS represents the parallax of the source. For
OGLE-2016-BLG-1227, the angular Einstein radius is mea-
sured from the obvious finite-source effects, but the microlens
parallax cannot be measured due to the short timescale of the
observed light curve, i.e., tE,1. We, therefore, estimate M and
DL by conducting Bayesian analysis of the event based on
the measured event timescale tE and the relative lens-source
proper motion µ. We use µ instead of θE because tE and θE
are highly correlated.
FIG. 8. -- Probability distributions of the lens mass of the planet host
(Mhost) and the lens-source separation (DLS) obtained from the Bayesian anal-
ysis. The solid curve curve is the distribution obtained with the combined θE
and tE constraint and the dotted curve is the distribution obtained with only
the tE constraint.
In the Bayesian analysis, we conduct a simulation of Galac-
tic lensing events using the prior models of the mass func-
tion of astronomical objects in the Galaxy and their physical
and dynamical distributions. For the mass function, we con-
sider both stellar and remnant lenses, i.e., black holes, neu-
tron stars, and white dwarfs, by adopting the Chabrier (2003)
model and the Gould (2000) model for the mass functions of
stars and remnants, respectively. In the simulation, lenses and
source are located following the physical distribution model
of Han & Gould (2003) and their motions are computed us-
ing the dynamical model of Han & Gould (1995). We pro-
duce 107 artificial lensing events, from which the probability
distributions of M and DL are obtained with the constraints of
the measured tE and µ.
In Figure 8, we present the probability distributions of
the lens mass of the host star (Mhost, upper panel) and the
lens-source separation (DLS, lower panel) obtained from the
Bayesian analysis. As indicated by the small angular Ein-
stein radius, the lens is estimated to lie close to the source,
and thus we present the distribution of DLS rather than DL.
To check the importance of the µ constraint, we present two
sets of distributions obtained with the combined µ and tE con-
straint (solid curves) and with only the tE constraint (dotted
curves). The distributions show that the lens mass estimated
with the additional µ constraint is substantially lower and the
lens-source separation is smaller than those estimated with the
single tE constraint. This indicates that the measured µ pro-
vides an important constraint on the physical lens parameters.
In Table 4, we list the estimated physical lens parameters.
We note that both the inner and outer 2L1S solutions result in
similar parameters, and thus we present the parameters based
on the inner 2L1S solution. The presented parameters are the
median values of the Bayesian distributions, and the upper
and lower limits correspond to the 15.9% and 84.1% of the
distributions. It is found that the lens is a planetary system
composed of a giant planet and a low-mass host star. The
Han et al.
8
Parameter
Mp (MJ)
Mhost (M⊙)
DLS (kpc)
a⊥ (au)
TABLE 4
PHYSICAL LENS PARAMETERS
tE + θE
0.79+1.30
−0.39
0.10+0.17
−0.05
1.21+0.96
−0.63
3.4+2.1
−1.0
Constraint
tE only
4.98+3.05
−2.94
0.68+0.42
−0.41
2.60+1.29
−1.13
11.5+3.1
−4.3
NOTE. -- The presented parameters are the median values of the
Bayesian distributions, and the upper and lower limits correspond to the
15.9% and 84.1% of the distributions.
masses of the planet and host are
and
Mp = 0.79+1.30
−0.39 MJ
Mhost = 0.10+0.17
−0.05 M⊙,
(9)
(10)
respectively. The planetary system is located in the bulge with
a line-of-sight separation from the source star of
DLS = 1.21+0.96
−0.63 kpc.
(11)
The planet and host are separated in projection by
a⊥ = 3.4+2.1
(12)
Considering that
the system is asl ∼
2.7 au(Mhost/M⊙) ∼ 0.4 au, the planet is a wide-separation
planet located well beyond the snowline of the host star.
−1.0 au.
the snowline of
7. DISCUSSION AND CONCLUSION
We analyzed the microlensing event OGLE-2016-BLG-
1227, for which the event timescale was short and the light
curve was affected by severe finite-source effects. The light
curve appeared to be that of a 1L1S event and the analysis
based on the 1L1S interpretation yielded a short timescale
and a very small angular Einstein radius, suggesting that the
lens could be an FFP. From the close inspection of the small
residuals from the 1L1S solution, we found that the residual
was explained by the existence of an additional widely sep-
arated heavier lens component, indicating that the lens was
a planetary system with a wide-separation planet rather than
an FFP. From the Bayesian analysis with the constraints of
the measured event timescale and relative lens-source proper
motion, we estimated that the lens was composed of a planet
with a mass Mp = 0.79+1.30
−0.39 MJ and a host star with a mass
Mhost = 0.10+0.17
−0.05 M⊙. It turned out that the planet was located
well beyond the snowline of the host with a projected sepa-
ration of a⊥ = 3.4+2.1
−1.0 au It was estimated that the lens was
located close to the source with a lens-source separation of
DLS = 1.21+0.96
The event demonstrates that detecting deviations from
1L1S light curves provides an important method to distin-
guish wide-separation planets from FFPs. Besides OGLE-
2016-BLG-1227, there were two planetary events, in which
planets were detected through isolated events and their widely
separated hosts were identified in lensing light curves. The
first case is MOA-bin-1 (Bennett et al. 2012). For this event,
the lensing light curve exhibited little lensing magnification
attributable to the host of the planet similar to OGLE-2016-
BLG-1227, but the planetary signal was entirely due to a brief
−0.63 au.
caustic feature. The second case is OGLE-2008-BLG-092
(Poleski et al. 2014). For this event, the planet was detected
through the isolated event channel, but in this case the host
of the planet was on the source trajectory, and gave rise to
a bump in the lensing light curve. OGLE-2016-BLG-1227
shows that there are a range of deviations in the signatures of
host stars and that it is possible to identify the existence of a
host even when a deviation is subtle.
Due to the unusual nature of OGLE-2016-BLG-1227, in
which the relative lens-source proper motion µ = θE/tE is well
determined, but the separate values of θE and tE are poorly
constrained, the information that can be obtained from high-
resolution follow-up observations would be different from
that of normal events.
If follow-up observations are con-
ducted to normal events with well estimated θE, the flux from
the host is measured and from this one can make a diagram
of the predicted host flux in M -- DL plane. Comparison of
this diagram to θE constraint in the same M -- DL plane will
allow one to determine M and DL from the intersection of
these two constraints, e.g., Yee (2015) and Fukui et al. (2019).
Even if θE is not known because of poor ρ measurement, the
event timescale tE is known. Then, from late time follow-up
imaging conducted when the source and lens are separated,
one can measure the lens-source separation ∆θ and therefore
the relative lens-source proper motion can be estimated by
µ = ∆θ/∆t, from which the angular Einstein radius is esti-
mated by θE = µtE . Here ∆t represents the difference between
the time of follow-up observation and t0.
For events with a well measured µ but with uncertain val-
ues of θE and tE, the time of follow-up observations can be
predicted. If follow-up observation is conducted using the Eu-
ropean Extremely Large Telescope (E-ELT) with an aperture
of 39 m, the full width half maxima (FWHM) in the J and
H band would be FWHM(J) ∼ 7.1 mas and FWHM(H) ∼
10.3 mas, respectively. Assuming that the lens and source can
be resolved when they are separated by ∼ 1.5× FWHM, the
required times for the resolution would be ∆t ∼ 13.5 years
and ∼ 19.6 years from J and H imaging observations, respec-
tively. These correspond to the years 2028 and 2035, respec-
tively. With a resolved host star, its distance DL and mass
Mhost would be constrained from the color and flux.
However, this does not necessarily imply that the planet
mass Mp = Mhost/q can also be well determined because the
mass ratio is poorly known. If one can estimate Mhost and DL
from the J and H color and magnitude, then there will two
If the lens is in the disk, one can estimate
possible cases.
πrel = au(D−1
S ), where DS ∼ 9 kpc. Then the Einstein
radius can be determined by the relation in Equation (1), al-
though uncertainty will be fairly large because Mhost and DL
are somewhat uncertain together with the uncertainty of the
source distance. If the lens is in the bulge, in contrast, it will
be difficult to estimate θE any better than from the microlens-
ing data. This will cause q and Mp to be poorly constrained.
L − D−1
Work by CH was supported by the grants of National
Research Foundation of Korea (2017R1A4A1015178 and
2019R1A2C2085965). Work by AG was supported by US
NSF grant AST-1516842 and by JPL grant 1500811. AG
received support from the European Research Council under
the European Union's Seventh Framework Programme (FP 7)
ERC Grant Agreement n. [32103]. The OGLE project has
received funding from the National Science Centre, Poland,
grant MAESTRO 2014/14/A/ST9/00121 to AU. This research
Wide-separation Microlensing Planet
9
has made use of the KMTNet system operated by the Korea
Astronomy and Space Science Institute (KASI) and the data
were obtained at three host sites of CTIO in Chile, SAAO
in South Africa, and SSO in Australia. We acknowledge the
high-speed internet service (KREONET) provided by Korea
Institute of Science and Technology Information (KISTI).
REFERENCES
Han, C., & Kang, Y. W. 2003, ApJ, 596, 1320
Han, C., Lee, C.-U., Udalski, A., et al. 2019, AAS, submitted
Kervella, P., Thévenin, F., Di Folco, E., & Ségransan, D. 2004, A&A, 426,
Alard, C., & Lupton, R. H. 1998, ApJ, 503, 325
Albrow, M. 2017, MichaelDAlbrow/pyDIA: Initial Release on Github, doi:
10.5281/zenodo.268049
Albrow, M., Horne, K., Bramich, D. M., et al. 2009, MNRAS, 397, 2099
Bennett, D. P., & Rhie, S. H. 2002, ApJ, 574, 985
Bennett, D. P., Sumi, T., Bond, I. A., et al. 2012, ApJ, 757, 119
Bensby, T., Yee, J. C., Feltzing, S., et al. 2013, A&A, 549, 147
Bessell, M. S., & Brett, J. M. 1988, PASP, 100, 1134
Chabrier, G. 2003, ApJ, 586, L133
Chung, S.-J., Han, C., Park, B.-G., et al. 2005, ApJ, 630, 535
Claret, A. 2000, A&A, 363, 1081
Fukui, A., Suzuki, D., Koshimoto, N. 2019, AAS, submitted
Gaudi, B. S., & Gould, A. 1997, ApJ, 486, 85
Gould, A. 1994, ApJ, 421, L71
Gould, A. 1997, ApJ, 480, 188
Gould, A. 2000, ApJ, 535, 928
Gould, A., Udalski, A., Monard, B., et al. 2009, ApJ, 698, L147
Gould, A., & Loeb, A. 1992, ApJ, 396, 104
Han, C. 2006a, ApJ, 638, 1080
Han, C. 2006b, ApJ, 644, 1232
Han, C., Chung, S.-J., Kim, D., et al. 2004, ApJ, 604, 372
Han, C., Gaudi, B. S., An, J. H., & Gould, A. 2005, ApJ, 618, 962
Han, C., & Gould, A. 1995, ApJ, 447, 53
Han, C., & Gould, A. 2003, ApJ, 592, 172
29
Kim, D.-J., Kim, H.-W., Hwang, K.-H., et al. 2018, AJ, 155, 76
Kim, S.-L., Lee, C.-U., Park, B.-G., et al. 2016, JKAS, 49, 37
Loeb, A., & Sasselov, D. 1995, ApJ, 449, L33
Mao, S., & Paczy´nski, B. 1991, ApJ, 374, L37
Mróz, P., Ryu, Y.-H., Skowron, J., et al. 2018, AJ, 155, 121
Mróz, P., Udalski, A., Bennett, D. P., et al. 2019, A&A, 622, A201
Nataf, D. M., Gould, A., Fouqué, P., et al. 2013, ApJ, 769, 88
Poleski, R., Skowron, J., Udalski, A., et al. 2014, ApJ, 795, 42
Refsdal, S. 1966, MNRAS, 134, 315
Valls-Gabaud, D. 1995, in Large Scale Structure in the Universe, ed.
J. P. Mücket, S. Gottlöber, & V. Müller (Singapore: World Scientific), 326
Udalski, A., Szyma´nski, M. K., & Szyma´nski, G. 2015, Acta Astron., 65, 1
Witt, H. J. 1995, ApJ, 449, 42
Witt, H. J., & Mao, S. 1994, ApJ, 430, 50
Wo´zniak, P. R. 2000, Acta Astron., 50, 42
Yee, J. C. 2015, ApJ, 814, L11
Yee, J. C., Shvartzvald, Y., Gal-Yam, A., et al. 2012, ApJ, 755, 102
Yoo, J., DePoy, D. L., Gal-Yam, A., et al. 2004, ApJ, 603, 139
|
1101.1103 | 1 | 1101 | 2011-01-05T21:29:37 | The rate of thermal atmospheric escape | [
"astro-ph.EP"
] | A formula is derived for the rate of thermal atmospheric escape, valid, and asymptotically exact, at low Knudsen number. | astro-ph.EP | astro-ph |
The rate of thermal atmospheric escape.
CCPP, Physics Department, New York University, 4 Washington Place, New York, NY 10003
Andrei Gruzinov
A formula is derived for the rate of thermal atmospheric escape, valid, and asymptotically
ABSTRACT
exact, at low Knudsen number.
1. The thermal escape problem
Consider a planetary atmosphere without stir-
ring and radiative heating or cooling above the
nominal surface. The atmosphere cannot be static
since a static atmosphere reaches thermodynamic
equilibrium. The Boltzmann distribution would
then give a finite density at infinite distances,
which is unphysical.
The atmosphere must be in a state of perma-
nent escape. We want to calculate the rate of es-
cape and the resulting temperature and density
profiles.
The thermal escape problem has a long history
(Jeans 1904, Parker 1958, Hunten 1982). But
in recent applications to planetary bodies, there
has been disagreement on which model – hydrody-
namic or free molecular flow – should be applied
(Johnson 2010). We believe that all issues have
been finally settled by a direct molecular dynam-
ics simulations of Volkov et. al. (2010). Here we
just show how the Volkov et. al. (2010) results
can be obtained analytically.
It would seem that the escape rate calculation
is only possible via molecular dynamics or Boltz-
mann numerical simulation, because the exobase
region, where the transition from hydrodynamic to
free molecular flow occurs, cannot be treated an-
alytically. But we show, for low Knudsen number
at the surface, that the exobase boundary condi-
tions are either irrelevant or can be deduced due
to large overlap between hydrodynamic and free
molecular flow regions. This allows one to calcu-
late the atmospheric escape using hydrodynamics
and to derive a formula for the escape rate.
For clarity, we approximate the molecular in-
1
0
-5
-10
-15
0
10
20
30
40
Fig. 1.- Escape rate for Kn = 10−8, cp = 5/2,
α = 1. Je < 2.5 is the Parker regime, 2.5 < Je <
24 is the Fourier regime, Je > 24 is the Jeans
regime.
teraction potential by a power law (this includes
hard-sphere gas as a limiting case). Then, by scal-
ing arguments, the escape rate must be given by
Ma = Mescape(Je, Kn).
(1)
Here we measure the escape rate by the radial ve-
locity at the surface v, and the velocity is repre-
sented by the isothermal Mach number
Ma ≡
The Jeans number is
.
v
pT /m
Je ≡
GM m
T R
,
(2)
(3)
where G is the gravitational constant, M is the
planet mass, m is the molecule mass, T is the sur-
face temperature, R is the surface radius. The
Knudsen number is Kn ∼ λ
R , where λ is the mean
free path. We define
Kn ≡
mκ
RρpT /m
,
(4)
where κ is the thermal conductivity, ρ is the den-
sity.
The escape rate formula is given in §2. The
formula includes three branches. In §3-5, we give
the recipes for calculating the branches. In §6, we
explain the origin of the recipes, describe numeri-
cal hydrodynamics which confirms the escape rate
formula, and compare our results to the molecular
dynamics simulations of Volkov et. al. (2010).
2. The escape rate formula
In the low-Knudsen limit,
Mescape =
MParker(Je), Je < cp;
MFourier(Je, Kn),
cp < Je . ln(Kn−1);
MJeans(Je), Je & ln(Kn−1).
(5)
Here cp is the heat capacity. The formula is
asymptotically exact (Appendix).
MParker is just the Parker wind (Parker 1958)
escape rate. It depends on Je and cp. For example,
for cp = 5/2,
MParker = p5/3.
(6)
2
Formula (5) limits the Je range over which the
Parker rate applies. The Parker regime is de-
scribed in §3.
MJeans is the Jeans (1904) escape rate. It de-
pends on Je only:
MJeans =
(Je + 1) exp(−Je)
√2π
(7)
Formula (5) extends the Kn range over which the
Jeans rate applies. The Jeans regime is described
in §4.
MFourier is proportional to Kn, with the Je-
dependent coefficient of proportionality:
MFourier(Je, Kn) = F (Je)Kn.
(8)
Here F (Je) depends on Je, cp, and
α ≡
For example, for α = 1,
d ln κ
d ln T
.
F (Je) =
1
Je − cp
(9)
(10)
For uniformity of notation, we call it the Fourier
regime. The Fourier regime is described in §5.
To illustrate the formula, Fig. 1 shows the es-
cape rate as a function of Je, for Kn = 10−8,
α = 1, and cp = 5/2. The different regimes were
extended to the intersection points. Fig.1 is in
qualitative agreement with molecular dynamics re-
sults of Volkov et. al. (2010). Quantitative com-
parison is done in §6, where we also show that
a simplified dissipative hydrodynamics model re-
produces the molecular dynamics results to a good
accuracy.
In the next three sections we calculate the es-
cape rate postulating the properties of the flow.
The assumptions made can be shown to be self-
consistent. But, in fact, what makes the results
trustworthy is that they do match the numerical
hydrodynamics results described in §6, and they
also match the molecular dynamics simulations of
Volkov et al (2010).
The Parker regime is a high-velocity ideal flow
with negligible thermal conductivity. The Fourier
regime is a flow with algebraically small velocity
and with thermal conduction balancing the ideal-
hydrodynamic energy flux. The Jeans regime is a
flow with exponentially small velocity, such that
thermal conduction dominates.
3. Parker regime
For Je < cp, Kn ≪ 1 one can use ideal hydro-
dynamics 1. To write equations in the simplest
form, we use scalings to set M = m = R = 1, and
also ρ = 1, T = 1, at the surface R = 1. Now
we need to calculate the outflow velocity at the
surface v0 = MParker, for any Je = G < cp.
We write stationary ideal hydrodynamics as
flux conservation, energy flux conservation, and
adiabaticity
ρvr2 = v0,
v2
2 + cpT − G
ρ = T cp
−1.
r = v2
0
2 + cp − G,
(11)
From eqs.(11)
v2
2
+ cp(cid:16) v0
r2v(cid:17)
1
cp−1
=
v2
0
2
+ cp − G +
G
r
.
(12)
This "algebraic" equation for v must have solu-
tions for all r > 1, such that the corresponding
density ρ = v0/(r2v) goes to zero at large r. This
requirement limits the range of possible values of
v0.
For cp < 3, and also for cp > 3, G <
2cp/(cp − 1), the smallest possible value is v0 =
pcp/(cp − 1),
in which case the sonic point is
at the surface.
For cp > 3 and larger G,
G > 2cp/(cp − 1), the sonic point moves out to
r > 1, and v0 decreases.
It is the smallest possible v0 that we call
MParker. Time-dependent ideal hydrodynamics
simulations and simulations of Volkov et al (2010)
confirm that the flow does approach the smallest
v0 allowed by stationary ideal hydrodynamics.
The above results can be written as
MParker = r cp
cp − 1
,
(13)
for cp < 3, and also for cp > 3, Je < 2cp/(cp − 1).
For cp > 3, 2cp/(cp − 1) < Je < cp, MParker de-
creases, reaching zero at Je = cp. The precise
value of MParker is then given implicitly by an
1 More precisely, dissipative effects must become negligible
after a possible boundary layer is passed. Whether the
boundary layer is present, depends on how the atmosphere
is fed. If the boundary layer is present, we simply move the
nominal surface up, beyond the boundary layer.
"algebraic" equation which can be derived from
eq.(12)2.
4. Jeans regime
Fix the Knudsen number and let the Jeans
number increase. Then the escape rate approaches
MJeans =
(Je + 1) exp(−Je)
√2π
(14)
– the standard Jeans rate – density normal-
ized flux of positive-energy positive-radial-velocity
molecules, calculated for Maxwellian distribution.
Eq.(14) is an exponentially accurate approxi-
mation (9% error at Je = 2.5, 0.008% error at
Je = 10) of the true escape rate for collision-
less molecular outflow with Maxwellian injection
at the surface, which in our notations is
0 uduR ∞
MJeans = R ∞
0 uduR ∞
R ∞
−∞ dvf v
−∞ dvf
.
(15)
Here v is the radial velocity, u is the absolute value
of the tangential velocity, f ∝ θ exp(−(u2 + v2)/2)
is the truncated Maxwellian. θ = 0 if simultane-
ously the energy is positive, (u2 + v2)/2 − Je > 0,
and the radial velocity v is negative; otherwise
θ = 1.
5. Fourier regime
Fix the Jeans number Je > cp and let the Knud-
sen number decrease. The escape rate then ap-
proaches
MFourier = F (Je)Kn.
(16)
The coefficient F (Je) is calculated below.
In the Fourier regime, the negative hydrody-
namic energy flux carried by the flow is almost
exactly compensated by the thermal conductivity:
r2κT ′ = Φ(cpT −
G
r
).
(17)
Here Φ = ρvr2 =const is the flux per steradian,
M = m = 1 was set by scaling, prime denotes
the r-derivative, kinetic energy and viscosity are
negligibly small.
Using scalings, we put T = 1 at r = 1, and then
solve eq.(17) for r > 1. We adjust Φ in such a way
2 (cid:16) Ma2
2 + c
p − Je(cid:17)cp
3
− 5
2 = (cid:16) cp
−1(cid:17)cp
cp
−1
(c
p −
5
2 )cp
− 5
2
4Ma
Je2 .
as to get the temperature going to zero at large
radii (otherwise the flow either stalls or becomes
singular at finite radii, never reaching an exobase,
§6). We want to solve the problem for arbitrary
α ≡ d ln κ/d ln T 3.
First consider just the case α = 1. One can
check numerically or analytically, that there is
only one solution of eq.(17) with correct asymp-
totic behavior at large r:
T =
1
r
Substitution into eq.(17) then gives
κ
T
= (G − cp)Φ.
Or, in different notations,
F (Je) =
1
Je − cp
.
(18)
(19)
(20)
For arbitrary α < 1, numerical integration of
eq.(17) shows that there is still a unique choice of
Φ which gives a non-singular zero-at-infinity tem-
perature. One can show that for 0 < Je − cp ≪ 1,
while for Je ≫ 1,
F =
1
Je − cp
.
F =
2
1 + α
1
Je
.
Joining the asymptotics as follows
(21)
(22)
(23)
.
(24)
F (Je) =
f (Je) ≈ 1 −
1
(Je − cp)f (Je)
Je − cp
1 − α
2
Je − 0.55cp
gives an error less than 0.3% for all relevant α, cp
and Je.
use stationary non-ideal hydrodynamics. The flux
conservation, the energy flux conservation, and the
momentum equations are
Φ = ρvr2,
Q = Φ( v2
2 + cpT − G
r ) − r2κT ′,
ρvv′ = −(ρT )′ − Gρ
r ))′.
(25)
Here Φ, Q are constant fluxes, η is the viscosity4.
The boundary conditions at r = 1 are
3 ηr2v(v′ − v
r3 (r3η(v′ − v
1
r ) − 4
r2 + 4
3
ρ = 1, T = 1, v = Φ
(26)
Given G = Je and κ(T = 1) = Kn, we must find
a pair (Φ, Q) that gives the "true" flow 5. Which
flows are "true" is decided by the boundary con-
ditions at the exobase, and it would appear that
one cannot proceed without a molecular dynamics
simulation of the exobase region.
But it turns out that for Kn ≪ 1 the mere
requirement that the exobase exists allows to cal-
culate the escape flow in the Parker and Fourier
regimes. In the Jeans regime, this trick no longer
works – there are many outflows with an exobase.
What allows to solve the problem in the Jeans
regime is a large overlap between hydrodynamic
and free molecular flow regions.
Parker regime – Take G < cp, and, as always,
κ = Kn ≪ 1.
If the initial velocity is large,
v0 ≡ Φ ∼ 1, the non-ideal hydrodynamics (25) de-
generates into the ideal hydrodynamics (11), and
we get the Parker escape.
Only for small velocities, v0 . κ, dissipative
effects can be significant. But numerical integra-
tion of (25) shows that for small v0 the exobase
is never achieved. Kn stays small at all radii, no
matter what energy flux Q one chooses.
What sets the boundary of the Parker regime
to Je = cp is the sign of the ideal hydrodynamic
energy flux in the small-velocity limit
Qh = Φ(cpT −
G
r
).
(27)
6. Hydrodynamics
To show that the escape rate formula (5) is
asymptotically exact at Kn → 0, and to improve
the accuracy in the non-asymptotic regime, we
3
α = 1 is a good approximation for real gases at relevant
temperatures; α = 0.5 for the hard-sphere gas, and we
want to reproduce the results of Volkov et. al. (2010)
4 We have assumed that the second viscosity vanishes. This
is true for the hard-sphere gas and allows precise compar-
ison to the results of Volkov et. al. For polyatomic gases,
the second viscosity must be included, but it turns out that
the viscosity effects are negligible at low Knudsen, unless
Je is very close to c
5 To solve (25) we also need v
′(1). But this boundary con-
′(1) would result in a bound-
dition is irrelevant. Wrong v
ary layer, and we have agreed to place the nominal surface
above the boundary layer.
p.
4
For Je < cp this is positive. Now we show that
this means no exobase for small v0.
For small v0, the density is approximately given
by the equilibrium condition
− (ρT )′ −
Gρ
r2 = 0.
(28)
At moderate G, the only way to reduce the density
enough to get an exobase, is by decreasing the
temperature at large distances. If the temperature
remains finite at infinity, the density saturates as
ρ ∝ exp(
G
T r
).
(29)
At small v0, we can drop the v2 terms from the
energy flux expression in (25), and get
r2κT ′ = Φ(cpT −
G
r
) − Q.
(30)
For the temperature to decrease, we must take a
positive flux Q. But finite Q then dominates the
r.h.s. of eq.(30) at large radii, and we get
This gives
r2κT ′ = −Q.
T ∝ r−
1
1+α .
(31)
(32)
This temperature decreases – but too slowly. For
α > 0, eqs.(28, 32) give a density growing at in-
finity.
In summary,
small-velocity flows have no
exobase, and large-velocity flows are ideal, giv-
ing the Parker flow. The lack of collisions beyond
the exobase cannot change the escape rate of the
highly supersonic outflow.
Fourier regime – At fixed Je > cp, as Kn de-
creases, the set of pairs (Φ, Q) which give solu-
tions with an exobase shrinks to a single point
(MFourier, 0). If (Φ, Q) 6= (MFourier, 0), the flow ei-
ther becomes singular at a finite radius or goes to
infinite radii, without ever reaching local Kn ∼ 1.
So in this case too, we can claim that MFourier is
an exact asymptotic even without knowing what
happens at the exobase.
The only way to reach an exobase is to set Q =
0 in eq.(30), and then tune Φ, as described in §5.
Jeans regime – Fix Kn and let Je increase.
The outflow velocity becomes so small that ther-
mal conduction in eq.(30) dominates and we get a
nearly constant temperature profile. Eq.(28) then
gives an exponential density. These temperature
and density are also predicted by the free molec-
ular outflow at high Je. We see that the two de-
scriptions agree.
We can therefore use hydrodynamic descrip-
tion well above the point r = Re where the
density scale height becomes comparable to the
mean free path, Kn(Re) = Je−1(Re). But for
Kn(r) ≫ Je−1(r), we can also use the free molec-
ular outflow results to calculate the fluxes (Φ, Q).
Now, we use hydrodynamics to integrate back to
r = 1. One can show, or just check numerically,
that this procedure, for Je ≫ ln Kn−1, gives the
Jeans rate (to leading order in Je in the prefactor).
Simplified hydrodynamics – Both the Fourier
and the Jeans regimes can be described simulta-
neously, in a much simplified version of hydrody-
namics which we give below. One can then hope
that even between the asymptotics this hydrody-
namic description will work (it does work for hard
spheres, see below).
The model has one adjustable parameter Kn0 ∼
1. We set T = 1, ρ = 1, Kn = Kn0 at r = 1, and
use the free molecular flow to calculate Φ and Q
at r = 1:
Φ =
e−G(G + 1)
√2π
, Q =
e−G((cp − 3
2 )G + cp − 1
2 )
√2π
.
(33)
We then integrate eq.(30) backwards in radius, to
r < 1. Simultaneously we integrate eq.(28) and
calculate Kn(r). The integration is terminated
when the desired value of Kn is reached. We ad-
just G so that the desired value of Je is reached
simultaneously with the desired value of Kn. This
gives the escape rate for these Kn and Je. One
also gets correct temperature and density profiles
for G ≫ 1 and also for r ≪ 1 at any G. We have
checked, by numerical integration of the full sys-
tem (25), that neglecting viscosity and inertia is
justified, again for G ≫ 1 and also for r ≪ 1 at
any G.
We have used the simplified hydrodynamics
(30,33) to establish the logarithmic accuracy of
the escape rate formula (5) in the Fourier-Jeans
transition region (Appendix).
Comparison with Volkov et. al. (2010) – Now
compare the predictions of our escape rate formula
eq.(5) to the molecular dynamics simulations of
5
Volkov et. al. (2010). We put cp = 5/2, α = 1/2.
We also remember that our definition of Kn is dif-
ferent by a constant factor from the conventional
hard-sphere Knudsen Knhs = 0.34Kn.
We find that for Je = 3, 10, 15, Knhs =
10−2, 10−3, 10−4
the simplified hydrodynamics
(30,33) (with Kn0 = 3) predicts the escape rates
which agree with the results of Volkov et.
al.
(2010) to about 10% for all points but one.
Let us discuss this discrepancy. For Je = 3,
the escape rate formula (5) gives for Knhs =
10−2, 10−3, 10−4 (using eqs.(16, 23, 24))
Ma
Knhs
= 6.4, 6.4, 6.4
(34)
The simplified hydrodynamics (33) gives
Ma
Knhs
= 5.6, 6.0, 6.2
(35)
the Parker wind model (§3). Both temperature
In the Jeans
and density decrease algebraically.
regime, we have a nearly isothermal slowly escap-
ing atmosphere, with exponential density. In the
Fourier regime, thermal conductivity balances the
ideal energy flow (§5). The temperature and the
density decrease algebraically.
I thank Robert Johnson for useful discussions.
Appendix
The escape rate formula (5) is asymptotically
exact in the following sense:
Je = const < cp, Kn → 0 :
Mescape
MParker → 1,
Je = const > cp, Kn → 0 :
(37)
(38)
(39)
Mescape
MFourier → 1,
Mescape
MJeans → 1,
Fig.4 of Volkov et al (2010) gives
Kn = const, Je → ∞ :
Ma
Knhs
= 5.0, 6.8, 4.1
(36)
The simplified hydrodynamics works as expected
– as we go deeper into the Fourier regime, pre-
dicted Ma approaches MFourier. The first two val-
ues of Volkov et. al. are also reasonable – the
error of MFourier decreases as we get deeper into
the Fourier regime. But the last point should have
been even closer to the Fourier value of 6.4. Some
50% error is unexpected.
We believe that it is the molecular dynamics
results at Je = 3, Knhs = 10−4 which have a large
error. We have checked, by integrating the full sys-
tem (25) numerically, that a hydrodynamic flow,
starting at Knhs = 10−4, Ma = 4.1× 10−4, Je = 3,
never goes above Knhs = 1.5 × 10−3 for any Q. It
seems unlikely that molecular dynamics can no-
ticeably deviate from hydrodynamics at such low
Knudsen numbers.
7. Summary
Formula (5) gives the thermal escape rate at
low Knudsen.
The escape occurs either in Parker, in Fourier,
or in the Jeans regime.
In the Parker regime,
we get a supersonic ideal hydrodynamical wind.
The temperature and density profiles are given by
ln(Mescape)
Je > cp, Kn → 0 :
ln(min(MFourier, MJeans)) → 1.
(40)
In words: (i) As Knudsen decreases at fixed
Jeans, the Parker and Fourier formulas approach
the true escape rate everywhere but near the
Parker-Fourier transition. As Knudsen decreases,
the Jeans number width of the Parker-Fourier
transition region shrinks to zero.
(ii) As Jeans
increases at fixed Knudsen, the Jeans formula ap-
proaches the true escape rate. (iii) In the Fourier-
Jeans transition interval, both formulas are valid,
but only with logarithmic accuracy.
REFERENCES
Jeans, J. H., 1904, "The dynamical theory of
gases", Chapter XVII, University Press, Cam-
bridge.
Parker, E. N., 1958, ApJ, 128, 664
Hunten, D. M., 1982, Planet. Space Sci., 30, 773
Johnson, R. E., 2010, ApJ, 716, 1573
Volkov, A. N., et al, 2010, arXiv:1009.5110
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
6
|
1902.09422 | 1 | 1902 | 2019-02-25T16:38:21 | Solar cycle variation in meteor rates | [
"astro-ph.EP"
] | Sixteen years of meteor radar data from the Canadian Meteor Orbit Radar (CMOR) were used to investigate the link between observed meteor rates and both solar and geomagnetic activity. Meteor rates were corrected for transmitter power and receiver noise, and seasonal effects were removed. A strong negative correlation is seen between solar activity, as measured with the 10.7 cm flux, and observed meteor rates. This lends support to the idea that heating in the atmosphere at times of elevated solar activity changes the scale height and therefore the length and maximum brightness of meteors; a larger scale height near solar maximum leads to longer, fainter meteors and therefore lower rates. A weaker negative correlation was observed with geomagnetic activity as measured with the $K$ index; this correlation was still present when solar activity effects were removed. Meteor activity at solar maximum is as much as 30\% lower than at solar minimum, strictly due to observing biases; geomagnetic activity usually affects meteor rates by less than 10 percent. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- 10 (2019)
Preprint 26 February 2019
Compiled using MNRAS LATEX style file v3.0
Solar cycle variation in radar meteor rates
M. D. Campbell-Brown,1(cid:63)
1Department of Physics and Astronomy, University of Western Ontario, London ON N6A 3K7 Canada
Accepted 2019 February 21. Received 2019 February 21; in original form 2019 January 10
ABSTRACT
Sixteen years of meteor radar data from the Canadian Meteor Orbit Radar
(CMOR) were used to investigate the link between observed meteor rates and both
solar and geomagnetic activity. Meteor rates were corrected for transmitter power and
receiver noise, and seasonal effects were removed. A strong negative correlation is seen
between solar activity, as measured with the 10.7 cm flux, and observed meteor rates.
This lends support to the idea that heating in the atmosphere at times of elevated solar
activity changes the scale height and therefore the length and maximum brightness of
meteors; a larger scale height near solar maximum leads to longer, fainter meteors and
therefore lower rates. A weaker negative correlation was observed with geomagnetic
activity as measured with the K index; this correlation was still present when solar
activity effects were removed. Meteor activity at solar maximum is as much as 30%
lower than at solar minimum, strictly due to observing biases; geomagnetic activity
usually affects meteor rates by less than 10 percent.
Key words: meteors -- Sun:activity
1 INTRODUCTION
The atmosphere is an excellent detector of meteoroids in the
millimeter size range; a single instrument can monitor hun-
dreds of square kilometres of at meteor heights. The effects
of the atmosphere cannot therefore be neglected, particu-
larly in the measurement of meteoroid flux.
Calculating meteoroid fluxes on the Earth involves
counting the number of meteors observed over a particu-
lar collecting area, and then correcting for observing biases.
Meteor patrol radars are particularly suited to calculating
fluxes, because they are not limited by weather or daylight
and can observe very large collecting areas over long peri-
ods of time. The number of echoes detected with the radar
depends on the flux of meteoroids, but also on the beam ge-
ometry, the transmitting power of the radar, the noise level
on the receivers, the efficiency of the detection software and
atmospheric effects.
1.1 Solar activity effects on meteor rates
The effects of the solar cycle on meteor rates were first stud-
ied by Bumba (1949), who reviewed visual meteor records
from 1844 to 1943, and found that the rates peaked 5 years
after maximum solar activity, and were lowest two years af-
ter solar minimum. However, it was not possible to rule out
(cid:63) E-mail: [email protected]
© 2019 The Authors
changes in cloud cover, for example, and he found similar
variations in the number of comets seen and meteorite falls
and finds, all of which seem unlikely to be affected by atmo-
spheric variations in the same way as meteors. The diversity
of records over the long time interval considered, in addition
to the fact that many of the meteor observations were at
the times of major showers, also imply that care should be
taken with the results.
The next careful analysis of meteor rates and solar ac-
tivity were done on data from the Omsala radar in Sweden,
which ran during the Perseids in 1953 and from 1956 to
1965, and also for a control interval in September in many of
those years. The transmitter power was lowered after 1959,
but an experiment was run alternating the power between
the two levels every hour in order to compare the rates, and
a correction factor was applied to the later data. Lindblad
(1967) found a dependence on sunspot numbers of the rates
of Perseids, with rates at solar minimum a factor of two
higher than at solar maximum. The control period, which
avoided the dates of major showers, showed a similar trend.
The data set included an anomalous increase in the meteor
rates in 1963, also observed in Ottawa, Canada (McIntosh
& Millman 1964) and Christchurch, New Zealand (Ellyett &
Keay 1964), which has been attributed to atmospheric heat-
ing from a volcanic eruption (Kennewell & Ellyett 1974). In
his initial work, Lindblad attributed the change in rates to a
simple change in atmospheric density; this, however, should
lead only to a raising or lowering of ablation heights, with
even the height ceiling moving up and down without chang-
2 M. D. Campbell-Brown
ing the number of observed meteors. In a later analysis, with
data up to 1972, Lindblad (1976) found the same trend, but
this time attributed the changing rates to a change in scale
height (density gradient); a smaller scale height and steeper
density gradient cause meteors to ablate over a shorter dis-
tance and reach a higher maximum brightness and therefore
electron density. During solar maximum, heating in the at-
mosphere causes to scale height to increase and meteors of
the same mass are fainter. As evidence of this change, Lind-
blad used simultaneous visual observations and radar ob-
servations of Perseids over the observing cycle to determine
ablation heights. He found that while the beginning heights
were the same over the solar cycle, the ending heights were
higher during solar minimum, which implies a smaller scale
height. There was a phase offset in the trend, with the lowest
meteor rates occurring one to two years before solar maxi-
mum and the highest rates 4 to 6 years after solar maximum.
Ellyett & Kennewell (1980) pointed out that, in fact, a
change in scale height has two effects: one is the change
in maximum brightness, and the other the shortening of
the trail. They showed that, if meteors are detected ran-
domly along their length, the rate of meteors will depend
on the mass distribution index of the population. If there
are more bright than faint meteors (a mass distribution in-
dex s less than 2), the rate of meteors will fall as the scale
height increases, opposite to the effect found by Lindblad.
If the mass index is 2, the two effects will cancel each other
out. For meteor populations with more faint meteors (s>2),
the observed rate will decrease with increasing scale height
(decreasing density gradient). The temperature of the atmo-
sphere also affects the meteor rate directly through the diffu-
sion rate; Elford (1980) looked at the theoretical rate of mass
loss as a function of height for an isothermal, winter and
summer temperature profile and found strong differences in
the trail length and maximum brightness of meteors. He pre-
dicted that this would change the rate of meteor echoes by
up to 70%, independent of actual changes of meteor flux on
the Earth; more echoes would be seen in the summer months
for each hemisphere. He suggested that there may also be a
diurnal effect.
Radar data from the Christchurch radar was exam-
ined by Prikryl (1983), who looked at the rates of mete-
ors within a few days of "enhanced" F10.7 flux (which is
strongly correlated with sunspot activity). He found a de-
crease of 10 to 15% at these times, except during solar
minimum, when there was no effect. Ellyett (1977) looked
at both the Springhill (Ottawa) and Christchurch radars,
and found a strong effect in the 1960 - 1961 dataset in
Christchurch; the effect was less during the same interval
at Springhill, which he attributes to seasonal effects, which
enhanced the solar cycle effects in the southern hemisphere
and partially cancelled them in the northern hemisphere.
There was a stronger negative correlation of sunspot num-
ber against meteor rates for the full 1958 -- 1966 Springhill
dataset. He looked at the correlation for daytime and night-
time echoes, and found them to be similar.
Up to the 1980s, it seemed clear that there was a nega-
tive correlation between solar activity (measured either with
number of sunspots or the flux at 10.7 cm) and meteor
rates, possibly with a delay of a few years. However, when
a much longer time series from 1958 to 1997 at the Ondre-
jov radar was analysed (Simek & Pecina 1999; Pecina &
Simek 1999; Simek & Pecina 2002), a positive correlation
was found between meteor rates and sunspot number, with
maximum sporadic rates occurring between 1 and 1.5 years
after peak solar activity. The echoes analysed were all over-
dense, with durations greater than 0.4 s, and slightly differ-
ent phase shifts for echoes lasting longer than 1 s. They at-
tributed previous negative correlations to small numbers (for
Bumba's visual study) and anomalously high meteor rates
in 1963, near solar minimum, for the Omsala, Springhill and
Christchurch radar studies.
A more recent analysis of visual data by Dubietis & Arlt
(2010) considered the rate of sporadic meteors, excluding the
antihelion source, in four time intervals through the year
free of major showers: January, March, July and September.
They required observations to have limiting magnitudes of
+5.5 or fainter, and averaged at least 10 observers for each
day for the years 1984 to 2006. They found that meteor rates
were highest just after maximum solar activity, with a time
offset of 1 to 2 years after solar maximum, very similar to
the findings of Simek and Pecina.
The most recent study of solar cycle effects on radar
meteor observations did not look at rates, but echo heights;
Stober et al. (2014) used data from the Canadian Meteor Or-
bit Radar (CMOR) 38 MHz system taken from 2002 to 2014,
and looked at the variation in the peak of the height dis-
tribution of observed echoes. In addition to a strong annual
variation, they found a change of 4 km between observations
taken during solar maximum and solar minimum, consistent
with a change in the neutral density of the atmosphere.
1.2 Geomagnetic effects
The effects of geomagnetic activity on meteor rates have
also been studied. Lindblad (1978) used the Omsala radar
dataset and looked particularly at the days around a sector
boundary passage by the Earth. He found that, following a
sector boundary passage, the global geomagnetic Cp index
increased and the meteor rate decreased, both with a delay
of 2 or 3 days. The effect seemed to be stronger in data
collected at night than data collected during the day.
The Springhill and Christchurch radar datasets were ex-
amined by Prikryl (1979, 1983), this time around the central
meridian passage of bright and faint green-corona regions on
the Sun. He also found an inverse correlation between the
Kp index and the rate of meteors with persistent echoes; in
this case, the Kp index reached a minimum and the rate of
meteors a maximum a few days after a bright central merid-
ian passage, and the reverse for faint green corona regions.
The Kp index is similar to the Cp index, except that Kp
varies between 0 and 9; and Cp varies from 0 to 2.5. The
effect was still present but less significant when echoes with
all durations were considered.
1.3 Current work
It is obvious that, particularly with respect to solar activ-
ity, the effects of the atmosphere on meteor rates are not
fully understood. This will affect, for example, precise mea-
surements of the flux of meteor showers from year to year.
In order to measure the long-term effects of the atmosphere
on meteor rates, it is necessary to have equipment which
MNRAS 000, 1 -- 10 (2019)
changes as little as possible over at least one solar cycle, and
to account for any changes carefully. In the present work,
the measured transmitter power and receiver noise of the
38 MHz system of the CMOR radar are used to correct the
observed rates, and the data are used to assess the effects of
solar and geomagnetic activity on the meteor rates.
2 OBSERVATIONS
The Canadian Meteor Orbit Radar (CMOR) consists of
three radars operating at 17.45, 29.85 and 38.15 MHz si-
multaneously. It is located in Tavistock, Ontario, Canada
(43.26N, 80.77W), and has been running reliably since 2002.
All three frequencies have a single, three-element yagi trans-
mitter and five two-element yagi receivers arranged in an
interferometer to locate echoes in the sky; the gain patterns
are effectively all-sky. The 29 MHz system has five remote
receivers to enable the calculation of trajectories and orbits.
This system has been upgraded to run at 15 kW peak power;
the other two systems at 17 and 38 MHz transmit approx-
imately 6 kW. The 17 MHz system suffers from noise and
terrestrial interference during the day. The 38 MHz system
detects the lowest rate of echoes, but it has been running
nearly continuously with no changes to its hardware or soft-
ware. This system is therefore best suited to studying long-
term trends.
The 38 MHz system has an effective limiting magnitude
of approximately +8, and overdense echoes are filtered out
by the detection software, leaving underdense and transition
echoes. It detects an average of 3200 echoes each day, for
a total of 16.7 million echoes in the period from January
2002 to October 2018. Of these, about 700,000 are on the
echo lines for major meteor showers during their most active
times; the others are sporadics or belong to minor showers.
The raw number of echoes for each day is shown in
figure 1. The annual variation is very clear, with a peak in
the summer months and the lowest rates in winter. There
are also obvious instrumental effects: higher rates for most
of 2007, lower rates in parts of 2002, and a time of low rates
in 2005, in addition to small outliers in other years.
2.1 Transmitter Power
One major influence on the rate of meteor echoes is the
power transmitted by the radar. In theory, the radar trans-
mits 6 kW continuously, but small changes in the equipment
can cause changes in the power. The power has been mea-
sured nearly continuously since 2002; from 2002 to 2011 an
in-line power meter which recorded the average power in
a text file was used, and since 2009, webcam images of a
digital meter were taken at 5 minute intervals. These mea-
surements were processed to find the average transmitted
power for each day, excluding times when the power was 0.
In some cases, this is because the power was sampled dur-
ing the cycling of the transmitter, which happens every 30
minutes; in other cases, the transmitter had been turned off
for maintenance.
Figure 2 shows the transmitter power as a function of
time. The power is stable for long stretches, but it does
change by about 25% over the course of sixteen years. In
MNRAS 000, 1 -- 10 (2019)
Solar cycle variation in meteor rates
3
particular, the power was high in 2007, when raw echo rates
were also unusually high.
A correction factor was applied to the meteor rates to
remove the effect of changing transmitter power; since the
amplitude of an echo is proportional to the square root of
the transmitted power, a factor of:
(cid:115)
CT x =
s−1
PT
PT r e f
was applied to the rates. The reference transmitter power
used was 5 kW, but the value chosen is not important, since
the goal is to change the rates to a common limiting mag-
nitude. The mass distribution index chosen was s = 2.1,
consistent with the value for sporadics measured for CMOR
data in Blaauw et al. (2011).
2.2 Receiver Noise
In general, the receiver noise on CMOR is dominated by cos-
mic sources, such as the galactic centre, the Sun and Cas-
siopeia A. However, there are instrumental and terrestrial
noise sources which occasionally dominate. The noise for
each day was calculated using the measured signal-to-noise
ratio (SNR) and amplitude of each echo.
Figure 3 shows the 38 MHz receiver noise in linear units.
The receiver noise was high in part of 2002 and 2005 (where
it accounts for anomalously low observed rates), as well as a
few days in other years. The rates were corrected in a similar
way to the transmitter power corrections, with:
(cid:115)
s−1
.
PNr e f
PN
CN =
The reference noise power was 330000, a typical value; again,
the value used does not affect the relative rates.
2.3 Signal-to-Noise Ratio
The higher noise in 2005 accounted for some of the drop in
rates, but even with the correction for noise level, some of
the rates are still anomalously low. The average signal-to-
noise ratio (in dB) for echoes was calculated for each day;
the results are shown in figure 4.
The typical value is approximately 12.5, but some days
have considerably higher averages; this was traced to bad
antenna phases, which meant that only very strong echoes
had reliable interferometry. No attempt was made to correct
the rates for this effect: a limit of 14 was chosen, and all days
with an average SNR greater than this were dropped from
the analysis. About 50 days were affected. Note that since
the average SNR was not used to correct the rates, it does
not matter whether linear or logarithmic units are used.
3 RESULTS
The raw rates for each day were corrected for transmitter
power and receiver noise. There remains the seasonal varia-
tion of the meteoroid environment, which must be removed
before a correlation can be sought between the rates and
4 M. D. Campbell-Brown
Figure 1. Raw number of echoes on 38 MHz system, 2002 to 2018. Solar longitude is zero on the first day of spring. Major showers are
marked.
Figure 2. Measured transmitter power for 38 MHz system, 2002 to 2018.
MNRAS 000, 1 -- 10 (2019)
Solar cycle variation in meteor rates
5
Figure 3. Measured receiver noise in arbitrary linear units for 38 MHz system, 2002 to 2018.
Figure 4. Average signal-to-noise ratio for 38 MHz system, in dB, 2002 to 2018.
solar and geomagnetic activity. The average corrected rate
was found for each degree solar longitude, and the rate for
each day was then divided by this number. This removes
any seasonally repeating changes in the meteor rate, includ-
ing changes in the sporadic sources and in showers which
repeat every year. The results, scaled to unity, are shown in
figure 5.
MNRAS 000, 1 -- 10 (2019)
Some deviations from the average are still visible, but
most of the obvious instrumental effects have been removed.
No showers remain visible in the corrected data.
6 M. D. Campbell-Brown
Figure 5. Echo rate relative to the average for each degree solar longitude, corrected for transmitter power and receiver noise, for 38
MHz system, 2002 to 2018.
3.1 Solar effects
The data can now be examined for atmospheric effects. The
flux of 10.7 cm radiation from the Sun is measured three
times a day1; these three values were averaged to get daily
numbers. The radar data covers an interval from just after
solar maximum in 2002, through a solar minimum in 2008
and 2009, to a second, lower maximum in 2014. Figure 6
shows the solar activity and relative corrected meteor rates
plotted against julian day. There is a suggestion that rates
are lowest at times of highest solar activity.
To quantify the effect of the solar cycle on meteor rates,
we can look at the correlation. Figure 7 shows the trend
between relative rate and solar activity; even with significant
scatter, there is a clear inverse dependence of rate on solar
activity. The R-squared of a linear fit is only 0.17. We tried
binning the data in 10-day intervals, to see if that would
reduce the noise and improve the fit, but the R-squared was
smaller for the binned data, 0.12. Considering earlier work,
we also introduced a phase offset for the solar activity data,
ranging up to 2 years: the results are shown in table 1. The fit
is slightly worse for offsets of up to one year, and noticeably
worse at 2 years.
(in units of 10−22 Wm−2Hz−1) as:
With zero phase, the activity varies with solar activity
Ncor =
N
−0.001205 ∗ F10.7 + 1.0759 .
(1)
Fluxes at solar maximum to are about 30% lower than at
solar minimum. Here, the correction is set so that at a F10.7
flux of 63×10−22 Wm−2 Hz−1, the smallest flux recorded in
1 http://www.spaceweather.gc.ca/solarflux/sx-en.php
Table 1. Goodness of fit of solar activity (F10.7 flux) and relative
meteor rate for phase offset in years
Phase (years) R-squared
0.00
0.25
0.50
0.75
1.00
1.25
1.50
1.75
2.00
0.17
0.13
0.15
0.15
0.11
0.10
0.12
0.10
0.06
our interval, has a correction of 1, and higher meteor rates
will be produced at higher solar fluxes.
To look at short-term effects on the atmosphere, the
activity of daytime and nighttime meteors were considered
separately. Meteors between 03 and 09 UT (10 pm to 4 am
local time) were used for the nighttime analysis, and from
15 to 21 UT (10 am to 4 pm local) were included in the
daytime analysis. All meteors had the same correction for
transmitter power, but the noise (which varies diurnally and
is 40% higher during the day) was separately calculated for
each subset of meteors.
The same analysis procedures were followed as in the
case of the full day dataset. The daytime echoes had a
very slightly higher correlation and steeper slope (slope =
-0.001586; intercept = 1.1609; r-squared = 0.18), while the
nighttime echoes had a lower correlation and shallower slope
(slope = -0.001111; intercept = 1.1117; r-squared = 0.1096).
MNRAS 000, 1 -- 10 (2019)
Solar cycle variation in meteor rates
7
Figure 6. Solar 10.7 cm flux (F10.7 index, in units of 10−22 Wm−2Hz−1) and relative meteor echo rate (corrected for transmitter power
and receiver noise) on 38 MHz system as a function of julian day.
Figure 7. Relative echo rate as a function of solar 10.7 cm flux (F10.7 index, in units of 10−22 Wm−2Hz−1) on 38 MHz system.
MNRAS 000, 1 -- 10 (2019)
8 M. D. Campbell-Brown
3.2 Geomagnetic effects
solar
The same relative meteor rates, corrected for instru-
mental effects and relative to the average for a par-
ticular
longitude, were used to look for varia-
tions due to geomagnetic effects. To maintain consis-
tency with previous work, we looked at the dependence
of rates on the global geomagnetic index Kp (ftp.gfz-
potsdam.de/pub/home/obs/kp-ap/), but we also used the
local K index from Ottawa, the closest geomagnetic observa-
tory (ftp.geolab.nrcan.gc.ca/pub/forecast/k indices/). The
unitless K index varies between 0 and 8, with two steps
between each, so the series [0o, 1+, 1-, 1o,...] was converted
to [0, 0.3, 0.7, 1.0,...].
There was a weak inverse correlation of meteor rates and
Local geomagnetic index. Phase offsets of 0 to 5 days were
tested, and the strongest correlation was found with zero
phase offset. Both the maximum effect and the correlation
were lower than for the solar activity variation case.
Solar activity and the geomagnetic index are not
strongly correlated, but the K index tends to be higher at
times of elevated solar activity (Fig. 8). To rule out solar ac-
tivity being responsible for the observed variation with ge-
omagnetic index, all the meteor rates were corrected using
the solar correction derived in the previous section (Eq. 1).
The results are shown in Fig. 9, with the line of best fit.
The r-squared value is only 0.028, indicating a very weak
correlation, but meteor rates are slightly lower when the
geomagnetic index is high. The slope is weaker than before
the data were corrected for solar activity (-0.015 compared
to -0.022), but a slight trend is still present. Even at high
K values, however, the straight line fit gives a correction of
only a few percent; the correction factor is:
Ncor =
N
−0.01535 ∗ K + 1.000 .
(2)
Here the lowest K index is 0, so the correction is 1 at K=0
and increases with increasing index, to a maximum of 16%
at the highest K of 9. Very few days of data are affected with
the highest K index, and most will have corrections less than
10%.
4 DISCUSSION
There is a clear anti-correlation between solar activity and
meteor rates on the 38 Mhz radar. The fact that there is
no delay between increasing solar activity and decreasing
meteor rates implies that the atmosphere is reacting rela-
tively quickly to solar heating. The fact that the decrease
in rates is more pronounced for meteors recorded during the
day could imply that density gradient changes occur over
short time intervals, but other factors might be responsible.
For example, if the ionosphere density and vertical extent
are larger at times of elevated solar activity, Faraday rota-
tion may be increased and meteor rates reduced during the
day because of the change in the polarization of the echoes.
This temporal effect should be investigated in any shower
analysis which looks at the variation of rates over periods of
less than a day.
The same analysis was performed on rates from the 29
MHz CMOR radar, using the measured power and noise
for that system. There was no trend of the rates on solar
activity, and visual inspection of the rates showed system-
atic increases and decreases in the rates which must be the
result of uncorrected changes in the radar sensitivity. The
29 MHz system has undergone many upgrades during its
lifetime, and the parameters used to correct the data here
are inadequate to describe the sensitivity changes. Assum-
ing the change in meteor rates with solar activity are caused
by a change in the scale height of the atmosphere, the rates
should change in the same way at any frequency. If some
other effect is involved, such as Faraday rotation, there may
be a wavelength dependence. Similarly, meteor brightness
should be affected in the same way as the maximum ion-
ization, and therefore optical rates should investigated for
solar effects. It may be more difficult, however, to correct
for changing observing conditions.
Meteor rates should certainly be corrected for solar
activity; the effects of geomagnetic activity are less pro-
nounced, but should be kept in mind. It is unclear how the
atmosphere at meteor heights reacts to different levels of
geomagnetic activity, and so whether the results here are
generally applicable.
5 CONCLUSIONS
Solar and geomagnetic activity affect the atmosphere, and
therefore the observed rate of meteors. Solar activity can
depress the observed meteor flux by 30% at the highest so-
lar flux, and should be taken into account when looking at
shower activity in different years. Geomagnetic activity has
a smaller effect, but should be considered in comparing rates
when the uncertainties are low.
The fact that the dependence of rates on solar activity
is slightly higher during the day implies that either the at-
mosphere responds very rapidly to changes in solar flux, or
that the dependence of meteor rates on solar activity is due
to some combination of atmospheric density changes and
short-lived effects, such as changes in the ionosphere. Fur-
ther observations at different frequencies could help to shed
light on this, and determine whether a diurnal correction
is dependent on frequency. The effects of Faraday rotation
due to changes in the ionosphere on meteor rates should
be further investigated by explicitly looking at the electron
density in the D-region; the role of sporadic E layer interfer-
ence, which is negligible at 38 MHz but very important at
lower frequencies, should also be examined.
When looking at long-term atmospheric effects, it is
very important that instrumental effects be carefully tracked
and corrected. This study emphasizes the need for long-term
monitoring with unchanging equipment, since corrections for
significant changes cannot always be successfully made.
ACKNOWLEDGEMENTS
Funding for this work was provided through NASA coop-
erative agreement 80NSSSC18M0046 and the Natural Sci-
ences and Engineering Research Council of Canada (Grant
no. RGPIN-2018-05474).
MNRAS 000, 1 -- 10 (2019)
Solar cycle variation in meteor rates
9
Figure 8. Local (OTT) geomagnetic index and solar activity from the 10.7 cm flux, in units of 10−22 Wm−2Hz−1.
Figure 9. Relative meteor activity, corrected for solar variation as well as instrumental and seasonal effects, as a function of Local
(OTT) geomagnetic index
MNRAS 000, 1 -- 10 (2019)
10 M. D. Campbell-Brown
REFERENCES
Blaauw R. C., Campbell-Brown M. D., Weryk R. J., 2011, MN-
RAS, 412, 2033
Bumba V., 1949, Bulletin of the Astronomical Institutes of
Czechoslovakia, 1, 93
Dubietis A., Arlt R., 2010, Earth Moon and Planets, 106, 105
Elford W. G., 1980, in Halliday I., McIntosh B. A., eds, IAU
Symposium Vol. 90, Solid Particles in the Solar System. pp
101 -- 104
Ellyett C., 1977, J. Geophys. Res., 82, 1455
Ellyett C. D., Keay C. S. L., 1964, Science, 146, 1458
Ellyett C. D., Kennewell J. A., 1980, Nature, 287, 521
Kennewell J. A., Ellyett C. D., 1974, Science, 186, 355
Lindblad B. A., 1967, Meddelanden fran Lunds Astronomiska Ob-
servatorium Serie I, 226, 1029
Lindblad B. A., 1976, Nature, 259, 99
Lindblad B. A., 1978, Nature, 273, 732
McIntosh B. A., Millman P. M., 1964, Science, 146, 1457
Pecina P., Simek M., 1999, A&A, 344, 991
Prikryl P., 1979, Bulletin of the Astronomical Institutes of
Czechoslovakia, 30, 321
Prikryl P., 1983, Bulletin of the Astronomical Institutes of
Czechoslovakia, 34, 44
Stober G., Matthias V., Brown P., Chau J. L., 2014, Geophys.
Res. Lett., 41, 6919
Simek M., Pecina P., 1999, in Baggaley W. J., Porubcan V., eds,
Meteoroids 1998. p. 87
Simek M., Pecina P., 2002, Earth Moon and Planets, 88, 115
This paper has been typeset from a TEX/LATEX file prepared by
the author.
MNRAS 000, 1 -- 10 (2019)
|
1710.00606 | 1 | 1710 | 2017-10-02T12:36:05 | Is There a Temperature Limit in Planet Formation at 1000 K? | [
"astro-ph.EP"
] | Dust drifting inward in protoplanetary disks is subject to increasing temperatures. In laboratory experiments, we tempered basaltic dust between 873 K and 1273 K and find that the dust grains change in size and composition. These modifications influence the outcome of self-consistent low speed aggregation experiments showing a transition temperature of 1000\,K. Dust tempered at lower temperatures grows to a maximum aggregate size of $2.02 \pm 0.06$ mm, which is $1.49 \pm 0.08$ times the value for dust tempered at higher temperatures. A similar size ratio of $1.75 \pm 0.16$ results for a different set of collision velocities. This transition temperature is in agreement with orbit temperatures deduced for observed extrasolar planets. Most terrestrial planets are observed at positions equivalent to less than 1000 K. Dust aggregation on the millimeter-scale at elevated temperatures might therefore be a key factor for terrestrial planet formation. | astro-ph.EP | astro-ph | Draft version June 12, 2018
Typeset using LATEX twocolumn style in AASTeX61
IS THERE A TEMPERATURE LIMIT IN PLANET FORMATION AT 1000 K?
Tunahan Demirci,1 Jens Teiser,1 Tobias Steinpilz,1 Joachim Landers,1, 2 Soma Salamon,1, 2 Heiko Wende,1, 2 and
Gerhard Wurm1
1Faculty of Physics, University of Duisburg-Essen, Lotharstr. 1, D-47057 Duisburg, Germany
2Center for Nanointegration Duisburg-Essen (CENIDE), University of Duisburg-Essen, Carl-Benz-Str. 199, D-47057 Duisburg, Germany
ABSTRACT
Dust drifting inward in protoplanetary disks is subject to increasing temperatures. In laboratory experiments, we
tempered basaltic dust between 873 K and 1273 K and find that the dust grains change in size and composition.
These modifications influence the outcome of self-consistent low speed aggregation experiments showing a transition
temperature of 1000 K. Dust tempered at lower temperatures grows to a maximum aggregate size of 2.02 ± 0.06 mm,
which is 1.49± 0.08 times the value for dust tempered at higher temperatures. A similar size ratio of 1.75± 0.16 results
for a different set of collision velocities. This transition temperature is in agreement with orbit temperatures deduced
for observed extrasolar planets. Most terrestrial planets are observed at positions equivalent to less than 1000 K. Dust
aggregation on the millimeter-scale at elevated temperatures might therefore be a key factor for terrestrial planet
formation.
Keywords: planets and satellites: formation - protoplanetary disks - astronomical databases: mis-
cellaneous
7
1
0
2
t
c
O
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
0
6
0
0
.
0
1
7
1
:
v
i
X
r
a
Corresponding author: Tunahan Demirci
[email protected]
2
Demirci et al.
1. INTRODUCTION
The number of observed extrasolar planets increases
continuously (Schneider et al. 2012, www.exoplanet.eu).
This now allows a statistical view on the distribution of
certain parameters. The semi major axes of gas plan-
ets can be rather small, placing them in close proximity
to their host stars (hot Jupiters). This is evidence of
migration as giants are supposed to be formed beyond
the water snowline (Kley 2000; Nelson & Benz 2003;
Alibert et al. 2005; Benz et al. 2014; Durmann & Kley
2015, 2017). For smaller Earth-like planets the situation
is different. Although detection by transits and radial
velocity works best for planets with small orbital peri-
ods, significantly fewer terrestrial planets are found in
the hot environment close to the star compared to the
cooler outer parts (exoplanet.eu 2017). This is shown
in Figure 1 with data taken from exoplanet.eu (2017).
However, we do not plot the radial distance here, instead
using the temperature attributed to the planet. For a
small planet this should essentially be the temperature
of radiative equilibrium with the stellar radiation.
Figure 1. Temperatures of observed planets with radii
smaller than 5 rEarth. Data are taken from exoplanet.eu
(2017). The 1000 K limit as well as the upper limit of the
condensation region at 1500 K are highlighted. The inset
summarizes the distribution of all shown planets over tem-
perature.
Overlaid are two lines of constant temperature. The
1000 K line corresponds to a change in aggregation ob-
served in our work (see below). 1500 K is approximately
the upper limit of the condensation region for typical
silicates. The database is still one with low statistics,
and detailed dependencies on temperature might be dis-
cussed. There is, e.g., a maximum at about 500 K, but
the distribution might just as well be considered flat in
a wide range. None of these statements are statistically
significant at the moment. Nevertheless, there is a clear
gap in the occurrence of planets between about 1000 K
and 1500 K.
For a planet forming at a larger distance from the star,
the inward drift would not need to stop at the 1000 K
line. Also, from a more general perspective of late planet
formation phases there is no obvious reason why planets
should not form, e.g. at 1100 K as likely as at 900 K.
That leaves the question whether this temperature gap
is inherited from the early phases of planet formation by
aggregation at elevated temperatures.
More often than not, planet formation and its limits
are considered in light of radial distance. It might be
argued that the lack of planets at higher temperatures
might be due to a closer distance and a related lack of
material for planet formation. Certainly, temperature
and distance are related, but we note that they are not
the same. Small changes in stellar mass already lead
to large luminosity and significant temperature varia-
tions. However, the local material available will not sig-
nificantly change. Also, it is currently often argued that
inner planetary systems are formed from material drift-
ing inward (Boley & Ford 2013; Wurm et al. 2013; Hu et
al. 2014). Inward drift is a classical mechanism for redis-
tribution of matter (Weidenschilling 1977). As a special
aspect, e.g., Ebel & Alexander (2011) suggested that
Mercury might have formed from dusty matter (chon-
dritic porous interplanetary particles) drifting inward
that were reprocessed in a region of higher temperatures
with a composition that is then very different from solar
composition. Considering such work, one motivation for
our experiments is a radial inward drift of a mix of ma-
terials (see below). The mass of material available for
planet formation does not necessarily depend only on
the radial distance to the star. We also note that cur-
rent planet temperatures and temperatures at the time
of dust aggregation might not be the same. They might
be comparable if the relevant aggregation occurred in a
passive disk. They might have been higher in a still ac-
tively accreting disk. However, accretion of dust at these
active times might have removed these aggregates again.
In any case, we do not claim to have the only answer to
planet formation in the inner disk at high temperature,
but we think it is curious that there is a temperature
limit in the exoplanet database that might be reason
enough to provide this somewhat different point of view
here and take it as motivation for our experimental work.
Collisional growth of micrometer dust to millimeter
aggregates is a first step in the process of planet for-
mation. Numerous experimental (Blum & Wurm 2008;
Guttler et al. 2010; Kruss et al. 2017) and theoretical
(Dominik & Tielens 1997; Wada et al. 2011) studies
showed that sticking collisions easily provide porous dust
012345050010001500200025003000planetaryradiusr[rEarth]temperatureT[K]exoplanet.eu(2017July10)5001000150020000123456T[K]countsaggregates in the millimeter range. Growth saturates at
this size though, as bouncing between particles becomes
dominant and further growth is stalled. This is there-
fore also known as the bouncing barrier (Zsom et al.
2010; Wada et al. 2011; Dr¸azkowska & Dullemond 2014;
Kelling et al. 2014; Kruss et al. 2016, 2017). Its existence
is a robust finding. Only the final size of the bouncing
agglomerates might be a "free" parameter depending on
the initial conditions as, e.g. the grain size distribution,
disk properties, collision velocities, or material.
In a number of current planet formation models, grav-
ity has to take over from here (Chiang & Youdin 2010).
This requires mechanisms to concentrate the bouncing
aggregates (Klahr & Henning 1997; Johansen et al. 2006;
Chiang & Youdin 2010). These are most efficient if dust
aggregates are larger than a certain minimum size (min-
imum Stokes number). Depending on the individual
model, the critical particle size varies from decimeter
down to millimeters. Recent findings set lower limits to
the minimum Stokes number than before (Yang et al.
2017). Larger dust agglomerates are favorable for all
of these models, as they can be concentrated more eas-
ily. However, it should be mentioned that there is also
a maximum Stokes number above which the streaming
instability does not seem to work. In a simplified view,
there are two important sizes at the transition from colli-
sional growth and concentration mechanisms. The max-
imum size of particles grown at the bouncing barrier and
the minimum size needed, e.g.
for streaming instabili-
ties do not largely overlap. In fact, it might rather be
a problem that they do not overlap at all (Dr¸azkowska
& Dullemond 2014).
It is therefore not a trivial de-
tail but instead is crucial how large aggregates can grow
via coagulation and even a small difference might decide
whether planetesimals locally form or not.
Considering a rather crude distinction in composition
and temperature, it is well known that the material has
a direct influence on collisional growth. Water ice be-
yond the (water) snowline, for example, is considered
to be more sticky than silicates (Dominik & Tielens
1997; Gundlach et al. 2011; Okuzumi et al. 2012; Au-
matell & Wurm 2013; Gundlach & Blum 2015; Musio-
lik et al. 2016b). Non-polar CO2 ice on the other side
behaves like silicates beyond its snowline (Musiolik et
al. 2016a). But these are only general trends. In this
scheme, silicates should have one behavior everywhere
beyond their "snowline", which is more commonly re-
ferred to as condensation region at about 1350− 1500 K
(Scott 2007). This might be questionable though.
In
particular for a mixture of refractory grains, the ac-
tual temperature even far below the sublimation might
be important. At high but not extreme temperatures
3
re-crystallization and (partial) sintering might already
change the mechanical properties of agglomerates as well
as the collision behavior of single grains. de Beule et al.
(2017) considered changes in (static) sticking properties
for tempered dust samples for the first time. Here, we
study the influence of the tempering temperature on the
bouncing limit of dust aggregates consisting of a mixture
of minerals (olivines, pyroxenes, iron oxides, ...).
2. GROWTH EXPERIMENTS
The basic experimental setup underlying this work
has been used in a number of earlier studies on dust
aggregation (Jankowski et al. 2012; Kelling et al. 2014;
Kruss et al. 2016, 2017) (Figure 2). The principle way
it works is as follows.
In a low-pressure environment
(here, p ≈ 500 Pa) capillaries on the order of the mean
free path of the gas molecules and heated on one end
efficiently pump gas from the cold to the warm side
(Knudsen 1909; Muntz et al. 2002, thermal creep). A
porous dust aggregate acts like a collection of capillaries
(de Beule et al. 2015; Koester et al. 2017; Steinpilz et al.
2017). Therefore, if dust aggregates are placed on a hot
surface at low ambient pressure, gas flows from the cold
top through the aggregate toward its bottom. This leads
to an overpressure (Knudsen 1909). If the overpressure
is large enough the dust aggregates are levitated on the
air cushion they produce. A large number of levitated
aggregates can be generated this way that then collide
with velocities in the mm s−1 to cm s−1 range.
This velocity is in the range expected for dust in pro-
toplanetary disks at the bouncing barrier, where aggre-
gates no longer grow by hit-and-stick but preferentially
bounce off each other (Guttler et al. 2010; Zsom et al.
2010). Kruss et al. (2017) showed that a bouncing bar-
rier evolves self-consistently if small ((cid:46) 100 µm) dust
particles are placed on the heater. How large the dust
grows at maximum has not been studied with this setup
before. For a given experimental setting (pressure p,
temperature T ) it will depend on the sticking properties
of the dust grains.
We use this setup here to test whether a dust sam-
ple that is subject to tempering for some time leads to
a different final outcome. A snapshot of a final aggre-
gate distribution after 15 minutes is shown in Figure 3.
Further growth is not observed for a longer duration.
Samples are always measured at about 900 K heater
temperature. Therefore, even if the tempering took
place at higher temperatures, the collisions were studied
when the samples were cooler again. This way we do not
trace any sticking properties related to viscosity effects,
instead only due to compositional variations. With the
given setup, we do not trace the phase transitions occur-
4
Demirci et al.
encountering other aggregates at velocities beyond their
fragmentation limit (vmax ≈ 0.2 m s−1). This reduces
the absolute size and shows that details of the setting
are important. However, by leaving all of parameters
the same for a certain study, the influence of tempering
can be traced by the maximum aggregate size. Here, the
important findings are relative changes of aggregate size
with the temperature used for heating the aggregates.
3. SAMPLES
The idea behind these experiments is that a dust sam-
ple that drifts inward within a protoplanetary disk is
subject to increasing temperatures, which changes the
composition and, related to this, the grain size and stick-
ing properties. In particular for dominating species like
pyroxenes and olivines, phase transitions already occur
at moderate temperatures of several hundred K, which
might be relevant (de Beule et al. 2017). How this re-
ally affects collisional growth has not been studied be-
fore. The basic requirement to test our hypothesis is
therefore a sample with silicates, and for convenience
we chose a basalt sample here.
3.1. Grain sizes
The basic sample is produced by milling a basalt sam-
ple to grain sizes in the micrometer range as expected for
dust in protoplanetary disks. From this sample parts are
used for tempering at different temperatures, and from
each subsample a part was used for detailed grain size
analysis by a particle sizer (Mastersizer 3000). For com-
parison we also used a quartz sample where we would
not expect any changes upon tempering and in collision
experiments. The volume size distribution puts empha-
sis on the grains dominating in mass and examples are
shown in Figure 4. Number size distributions are shown
in Figure 5.
The number-dominating as well as the volume-
dominating grain sizes continuously increase with tem-
perature for the basalt samples. Possible causes for
this include compositional changes (see below). Sinter-
ing could also be a reason for the observed increase in
the grain sizes. The 120 h basalt sample heated at a
temperature below 1000 K shows no significant change
compared to the 1 h sample. For the sample heated
above 1000 K the grain size is somewhat larger com-
pared to the 1 hour sample. This indicates that time
has an influence here. However, this would only support
our general findings on collisional evolution as given be-
low. As expected, we can see no significant changes
for the pure quartz samples. Tab. 1 summarizes the
most likely grain sizes λmax and the mean grain sizes
Figure 2. Laboratory levitation setup to study the aggre-
gation of dust at low velocities (Kruss et al. 2016, 2017).
Figure 3. Snapshot of the initial dust placed on the heater
and the final size of aggregates after 15 minutes of growth.
The initial sample was sieved onto the heater through a
125 µm mesh.
ring at 700 or 800 K as studied by de Beule et al. (2017),
rather we trace changes at higher temperatures.
We used two slightly different settings for the tem-
perature of the heater, which result in different mean
mean ≈ 2 cm s−1 for the first and
translation velocities (v(1)
mean ≈ 1 cm s−1 for the second heater setting) of the ag-
v(2)
gregates and thus also in different maximum aggregate
sizes. The difference is due to the largest aggregates
heater &light sourcediffusorconcave lenscameraaggregates1 mmTable 1. Most Likely Grain Size λmax and Mean Grain Size λmean
of the Used Samples in Dependence of the Tempering Tempera-
ture and Tempering Duration.
5
Material Temperature Duration λmax
a
λmean
b
basalt
basalt
basalt
basalt
basalt
basalt
basalt
basalt
basalt
basalt
basalt
quartz
quartz
quartz
(K)
873
973
998
1023
1073
1123
1173
1223
1273
973
1073
873
1073
1273
(hour)
(µm)
1
1
1
1
1
1
1
1
1
120
120
1
1
1
0.588
0.616
0.616
0.620
0.629
0.641
0.654
0.687
0.697
0.622
0.64
-
-
-
(µm)
3.5 ± 0.3
6.1 ± 0.6
3.4 ± 0.3
4.2 ± 0.5
5.6 ± 0.7
5.2 ± 0.5
8.2 ± 0.9
10.7 ± 1.2
18.9 ± 4.8
5.7 ± 0.6
5.1 ± 0.5
3.3 ± 0.3
3.1 ± 0.2
3.2 ± 0.2
aDetermined from the number size
distributions (see Figure 5). Tolerance ∆λmax = 0.006 µm.
b Determined from the volume size distributions (see Figure 4)
3.2. Composition
To investigate possible changes in sample composition,
Mossbauer spectra of basaltic dust heated to 873-1073 K
for 1 h were recorded in transmission geometry and con-
stant acceleration mode at 80 K using a l-He bath cryo-
stat.
The spectrum of material heated to 873 K shown in
the top panel of Figure 6 displays a complex spectral
structure containing several subspectra. Two domi-
nant doublet subspectra with high isomer shift δ and
quadrupole splitting ∆EQ are visible, indicating Fe2+-
bearing paramagnetic minerals. Absolute numbers of
δ ≈ 1.25 mm s−1 (relative to α-Fe at ambient tempera-
ture) and ∆EQ of ca. 3.0 mm s−1 and 2.3 mm s−1, re-
spectively, are consistent with literature values reported
for iron-bearing minerals of the olivine (green) and py-
roxene (blue) group (Oshtrakh et al. 2007). Addition-
ally, the spectrum contains a minor doublet of smaller
quadrupole splitting and isomer shift (orange) and a
broad asymmetric sextet (dark red), whose hyperfine
parameters point toward ferric oxide. This combination
Figure 4. Volume size distributions of dust grains in the
different samples; the distributions were measured by a com-
mercial instrument based on light scattering (Malvern Mas-
tersizer 3000).
Figure 5. Number size distributions of dust grains in the
different samples (Malvern Mastersizer 3000). The inset
shows the most likely dust grain diameter λmax for the dif-
ferent heating temperatures and heating times.
λmean in dependence of the tempering temperature and
tempering duration.
basaltsample873K1073K1273K12345671/VtotaldV/dλ[μm-1]quartzsample873K1073K1273K0.515105010001234567diameterλ[μm]1/VtotaldV/dλ[μm-1]basaltsample873K1073K1273K0.51.01.52.005101520diameterλ[μm]1/NtotaldN/dλ[μm-1]λmaxat1273K●●●●●●●●●■■heatingtime●1h■120h9001000110012000.600.620.640.660.680.70temperatureT[K]λmax[μm]6
Demirci et al.
implies the presence of nanophase ferric oxide with par-
tial superparamagnetic properties, although the pres-
ence of additional ferric paramagnetic minerals cannot
be ruled out (Morris et al. 1993).
Figure 7. Relative spectral areas of identified mineral sub-
spectra in Mossbauer spectra of heated basaltic dust.
above the Morin transition, which is known to be sup-
pressed in hematite nanocrystals ( Ozdemir et al. 2008).
Thereby, it may indicate a phase transition from ferri-
magnetic magnetite and/or maghemite (dark red) to an-
tiferromagnetic hematite nanocrystals (gray). The de-
crease in line width, on the other hand, could originate
from structural ordering or, alternatively, from a moder-
ate increase in hematite crystal size, resulting in a minor
variation of superparamagnetic relaxation.
4. BOUNCING BARRIER SIZE
Figure 6. Mossbauer spectra of basaltic dust heated at
873-1073 K recorded at 80 K.
Spectra of basaltic dust heated to higher tempera-
tures display widely similar spectral structures. How-
ever, the spectral areas of olivine and pyroxene decrease
upon heating to the benefit of superparamagnetic as well
as magnetically blocked ferric oxide. Relative spectral
areas of the identified mineral subspectra are displayed
comparatively in Figure 7 for different heating temper-
atures to allow a better understanding of phase tran-
sitions and possible structural changes of the basaltic
material upon heating. Presumably, the decrease in
olivine and pyroxene spectral area corresponds to a par-
tial phase transition of Fe2+-bearing silicates to iron
oxide, which takes place at temperatures up to about
1000 K. A partial phase transition refers to the fact that
the transition does not occur for the entire material.
Upon further inspection, a continuous decrease in
sextet line width is evident, as well as a change of
the sextet quadrupole level shift from ∼ 0 mm s−1 to
−0.2 mm s−1. This value is characteristic for hematite
(cid:113) A
Figure 8 and 9 show the maximum aggregate sizes
that formed for the different dust samples tempered at
different temperatures. The maximum aggregate size
π is determined by the area A of the two-
d = 2
dimensional projection of the aggregate.
mean ≈ 2 cm s−1. For v(2)
There is a clear offset in growth at 1000 K. Dust tem-
pered at lower temperatures grows larger than the dust
processed at higher temperatures. This is true for both
heater settings. The ratio of the average diameter below
1000 K to the value above this temperature is 1.75±0.16
mean ≈ 1 cm s−1 this size ra-
for v(1)
tio is 1.49 ± 0.08. For comparison we also measured
the largest aggregate size for a pure quartz dust sample
(see Figure 9). These quartz results confirm that with-
out changes in composition we do not see any change in
growth. It also implies that the jump in aggregate size is
obviously not just an artifact by the thermal processing
and sample handling.
It is highly probable that internal phase transitions
are responsible for the change in growth. Grain size in-
crease might be a reason for less sticking. However, the
number-dominating size does not change spontaneously
8408401050-10-50510840pyroxenemagnetite/maghemiteolivinehematitesuperparamagneticiron oxiderelative absorption ∆I/I0[%]873K973K1023K1073Kvelocity v[mm/s]900950100010500102030405060superparamagn. Fe-oxidemagn. blocked Fe-oxideolivinepyroxene7
(see sec. 3.2). This could change the grain surface, but
without further investigation, we can neither confirm
nor exclude this presumption. This does not devalue
the present study though. It is a fact that the growth
changes with temperature whatever the specific reason.
5. CONCLUSION
Our results show that aggregate growth in protoplan-
etary disks might depend on the temperature history. A
certain (large) size might be a prerequisite for efficient
concentration mechanisms like the streaming instability
(Youdin & Goodman 2005; Johansen et al. 2006).
If
so, larger aggregate growth might promote the local for-
mation of planetesimals "outside of 1000 K". Certainly,
modifications of orbits occur at later times but from a
probabilistic point of view there might be some memory
left from the planetesimal formation process and this
indeed seems to correlate to the exoplanet database.
We see a strong correlation between the temperature
of a dust sample and the maximum size of aggregates
formed until they reach the bouncing barrier. There is
a clear separation in favored growth below 1000 K and
suppressed growth beyond 1000 K. The same correlation
is found between local temperature and the occurrence
of extrasolar terrestrial planets, which seem to be sparse
at temperatures higher than 1000 K. There are still some
steps in between millimeter dust and planets. However,
planetesimal formation, e.g., triggered by streaming in-
stabilities, would favor larger aggregates, and terrestrial
planets might form locally. Assuming this, the almost
too-perfect correlation between early dust growth and
planet locations is tantalizing. As a caveat, we only
considered a special dust sample in the experiments
and, e.g. a heater in air, so this is certainly still far
from optimized. Also, the local temperatures might vary
somewhat at the times planetesimals formed. Taken to-
gether, future experiments need to show whether 1000 K
is a general mark showing a transition in collision ex-
periments for a wider range of mineral compositions of
dust. However, our results already strongly indicate that
the growth of millimeter-size dust at high temperatures
might be a key factor in the formation of terrestrial plan-
ets in the inner regions of protoplanetary disks.
ACKNOWLEDGEMENTS
This project is funded by DFG grant KE 1897/1-
1, SPP 1681 grant WE2623/7-2, and FOR 1509 grant
WE2623/13. We also thank the referee for a construc-
tive review of the paper.
Figure 8. Maximum size of aggregates grown for samples
tempered at different temperatures for heater setting 1. The
dashed lines mark the average values below and above the
1000 K value that seems to divide two regimes. The aver-
age value below 1000 K is 1.75 ± 0.16 times the value above
1000 K. The basalt samples heated for 120 h do not differ
within the accuracy from the 1 h samples. The displayed
data points are average values of 49 measurements.
Figure 9. Same as Figure 8, but for different heater tem-
perature (basalt sample). The average value below 1000 K
is 1.49 ± 0.08 times the value above 1000 K. In addition, re-
sults from the quartz samples are shown. The displayed data
points are average values of 69 measurements.
across the transition, rather it increases continuously.
We also do not see a sudden transition at tempera-
tures near 1000 K in the mean grain size for the volume-
dominating grains. Even the Mossbauer spectra do not
show a sudden change near this temperature. We there-
fore cannot pin down one special reason why sticking
changes at a certain temperature. de Beule et al. (2017)
saw strong morphological changes on the grains of their
samples at lower temperatures. Whether grain morphol-
ogy is also important for the smaller grains studied here
is up to future work. The Mossbauer spectra may in-
dicate a phase transition and/or a structural ordering
■■■■■■◆◆basalt(vmean≈2cm/s)■theating=1h◆theating=120h90010001100120013000.60.81.01.21.41.61.8temperatureT[K]diameterd[mm]■■■■■■■●●●theating=1h■basalt,vmean≈1cm/s●quartz,vmean≈2cm/s90010001100120013000.51.01.52.0temperatureT[K]diameterd[mm]8
Demirci et al.
REFERENCES
Alibert, Y., Mordasini, C., Benz, W., & Winisdoerffer, C.
2005, A&A, 434, 343
Aumatell, G., & Wurm, G. 2013, MNRAS, 437, 690
Benz, W., Ida, S., Alibert, Y., Lin, D., & Mordasini, C.
2014, in Protostars & Planets VI, ed. H. Beuther et al.
(Tucson, AZ: Univ. Arizona Press), 691
Blum, J., & Wurm, G. 2008, ARA&A, 46, 21
Boley, A. C., & Ford, E. B. 2013, arXiv, 1306.0566
Chiang, E., & Youdin, A. N. 2010, AREPS, 38, 493
de Beule, C., Wurm, G., Kelling, T., Koester, M., &
Kocifaj, M. 2015, Icar, 260, 23
de Beule, C., Landers, J., Salamon, S., Wende, H., &
Wurm, G. 2017, ApJ, 837, 59
Dominik, C., & Tielens, A. G. G. M. 1997, ApJ, 480, 647
Dr¸azkowska, J., & Dullemond, C. P. 2014, A&A, 572, A78
Durmann, C., & Kley, W. 2015, A&A, 574, A52
Durmann, C., & Kley, W. 2017, A&A, 598, A80
Ebel, D. S., & Alexander, C. M. O' D. 2011, P&SS, 59, 1888
exoplanet.eu, "The Extrasolar Planets Encyclopedia", 2017
July 10 (www.exoplanet.eu)
Klahr, H., & Henning, T. 1997, Icar, 128, 213
Kley, W. 2000, MNRAS, 313, L47
Knudsen, M. 1909, AnPhy, 336, 633
Koester, M., Kelling, T., Teiser, J., & Wurm, G. 2017,
Ap&SS, 362, 171
Kruss, M., Demirci, T., Koester, M., Kelling, T., & Wurm,
G. 2016, ApJ, 827, 110
Kruss, M., Teiser, J., & Wurm, G. 2017, A&A, 600, A103
Morris, R. V., Golden, D. C., Bell III, H. V. Lauer Jr.,
R. B. & Adams, J. B. 1993 GeCoA, 57, 4597
Muntz, E. P., Sone, Y., Aoki, K., Vargo, S., & Young, M.
2002, JVSTA, 20, 214
Musiolik, G., Teiser, J., Jankowski, T., & Wurm, G. 2016,
ApJ, 818, 16
Musiolik, G., Teiser, J., Jankowski, T., & Wurm, G. 2016,
ApJ, 827, 63
Nelson, A. F., & Benz, W. 2003, ApJ, 589, 556
Okuzumi, S., Tanaka, H., Kobayashi, H., & Wada, K. 2012,
ApJ, 752, 106
Oshtrakh, M. I., Petrova, E. V., Grokhovsky, V. I., &
Semionkin, V. A. 2007, HyInt 177, 65
Gundlach, B., Kilias, S., Beitz, E., & Blum, J. 2011, Icar,
Ozdemir, O, Dunlop, D. C., & Berqu´o, T. S. 2008, GGG, 9,
214, 717
Q10Z01
Gundlach, B., & Blum, J. 2015, ApJ, 798, 34
Schneider, J., Le Sidaner, P., Savalle, R., & Zolotukhin, I.
Guttler, C., Blum, J., Zsom, A., Ormel, C. W., &
2012, adass XXI, 461, 447
Dullemond, C. P. 2010, A&A, 513, A56
Hayashi, C., Nakazawa, K., & Nakagawa, Y. 1985,
Protostars and Planets II (Tucson, AZ: Univ. Arizona
Press), 1100
Hu, X., Tan, J. C, & Chatterjee, S. 2014, in IAU Symp.
310, Complex Planetary Systems, ed. Z. Knezevi´c & A.
Lemaıtre (Cambridge: Cambridge Univ. Press), 66
Jankowski, T., Wurm, G., Kelling, T., Teiser, J., Sabolo,
W., Guti´errez, P. J., & Bertini, I. 2012, A&A, 542, 80
Johansen, A., Klahr, H., & Henning, Th. 2006, ApJ, 636,
1134
Kelling, T., Wurm, G., & Koester, M. 2014, ApJ, 783, 111
Scott, E. R. D., 2007, AREPS, 35, 577
Steinpilz, T., Teiser, J., Koester, M., Schywek, M., &
Wurm, G. 2017, MiST, 29, 235
Wada, K., Tanaka, H., Suyama, T., Kimura, H., &
Yamamoto, T. 2011, ApJ, 737, 36
Weidenschilling, S. J. 1977, MNRAS, 180, 57
Wurm, G., Trieloff, M., & Rauer, H. 2013, ApJ, 769, 78
Yang, C. C., Johansen, A., & Carrera, D., 2017, A&A, in
press (arXiv:1611. 07014)
Youdin, A. N., & Goodman, J. 2005, ApJ, 620, 459
Zsom, A., Ormel, C.W., Guttler, C., Blum, J., &
Dullemond, C. P. 2010, A&A, 513, A57
|
1911.00609 | 1 | 1911 | 2019-11-01T22:59:51 | Observational Investigation of the 2013 Near-Earth Encounter by Asteroid (367943) Duende | [
"astro-ph.EP"
] | On 15 February 2013, the asteroid 367943 Duende (2012 DA14) experienced a near-Earth encounter at an altitude of 27,700 km or 4.2 Earth radii. We present here the results of an extensive, multi-observatory campaign designed to probe for spectral and/or rotational changes to Duende due to gravitational interactions with the Earth during the flyby. Our spectral data reveal no changes within the systematic uncertainties of the data. Post-flyby lightcurve photometry places strong constraints on the rotation state of Duende, showing that it is in non-principal axis rotation with fundamental periods of P_1 = 8.71 +/- 0.03 and P_2 = 23.7 +/- 0.2 hours. Multiple lightcurve analysis techniques, coupled with theoretical considerations and delay-doppler radar imaging, allows us to assign these periods to specific rotational axes of the body. In particular we suggest that Duende is now in a non-principal, short axis mode rotation state with a precessional period equal to P_1 and oscillation about the symmetry axis at a rate equal to P_2. Temporal and signal-to-noise limitations inherent to the pre-flyby photometric dataset make it difficult to definitively diagnose whether these periods represent a change imparted due to gravitational torques during the flyby. However, based on multiple analysis techniques and a number of plausibility arguments, we suggest that Duende experienced a rotational change during the planetary encounter with an increase in its precessional rotation period. Our preferred interpretation of the available data is that the precession rate increased from 8.4 hours prior to the flyby to 8.7 hours afterwards. A companion paper by Benson et al. (2019) provides a more detailed dynamical analysis of this event and compares the data to synthetic lightcurves computed from a simple shape model of Duende. (abbreviated abstract) | astro-ph.EP | astro-ph |
Observational Investigation of the 2013
Near-Earth Encounter by Asteroid (367943)
Duende
Nicholas A. Moskovitz a, Conor James Benson b,
Daniel Scheeres b, Thomas Endicott c, David Polishook d,
Richard Binzel e, Francesca DeMeo e, William Ryan f,
Eileen Ryan f, Mark Willman g, Carl Hergenrother h,
Arie Verveer i, Tim Lister j, Peter Birtwhistle k,
Amanda Sickafoose (cid:96),e,o, Takahiro Nagayama m, Alan Gilmore n,
Pam Kilmartin n, Susan Bennechi o, Scott Sheppard p,
Franck Marchis q, Thomas Augusteijn r, Olesja Smirnova r
aLowell Observatory, 1400 West Mars Hill Road, Flagstaff, AZ 86001 (U.S.A)
bUniversity of Colorado Boulder, Boulder, CO 80305
cUniversity of Massachusetts Boston, Boston, MA 02125
dWeizmann Institute of Science, 234 Herzl St. Rehovot 7610001, Israel
eMassachusetts Institute of Technology, EAPS, Cambridge, MA 02139
f New Mexico Institute of Mining and Technology, Socorro, NM 87801
gInstitute for Astronomy, University of Hawaii, Hilo, HI 96720
hUniversity of Arizona, LPL, Tucson, AZ 85721
iPerth Observatory, Bickley, WA 6076 (Australia)
Preprint submitted to Icarus
5 November 2019
jLas Cumbres Observatory Global Telescope Network, Goleta, CA 93117
kGreat Shefford Observatory, Berkshire (England)
(cid:96)South African Astronomical Observatory (South Africa)
mKagoshima University (Japan)
nUniversity of Canterbury, Mt John University Observatory, Lake Tekapo (New
Zealand)
oPlanetary Science Institute, 1700 E. Fort Lowell Rd., Tucson, AZ 85719
pDepartment of Terrestrial Magnetism, Carnegie Institution for Science,
Washington, DC 20015
qSETI Institute, Carl Sagan Center, 189 Bernardo Ave., Mountain View CA
94043
rNordic Optical Telescope, La Palma (Spain)
Copyright c(cid:13) 2019 Nicholas A. Moskovitz
Number of pages: 50
Number of tables: 5
Number of figures: 12
2
Proposed Running Head:
The 2013 Near-Earth Encounter of Asteroid Duende
Please send Editorial Correspondence to:
Nicholas A. Moskovitz
Lowell Observatory 1400 West Mars Hill Road
Flagstaff, AZ 86001, USA.
Email: [email protected]
Phone: (928) 779-5468
3
ABSTRACT
On 15 February 2013, the asteroid 367943 Duende (2012 DA14) experienced a
near-Earth encounter at an altitude of 27,700 km or 4.2 Earth radii. We present
here the results of an extensive, multi-observatory campaign designed to probe
for spectral and/or rotational changes to Duende due to gravitational inter-
actions with the Earth during the flyby. Our spectral data reveal no changes
within the systematic uncertainties of the data. Post-flyby lightcurve photom-
etry places strong constraints on the rotation state of Duende, showing that it
is in non-principal axis rotation with fundamental periods of P1 = 8.71± 0.03
and P2 = 23.7 ± 0.2 hours. Multiple lightcurve analysis techniques, coupled
with theoretical considerations and delay-doppler radar imaging, allows us to
assign these periods to specific rotational axes of the body. In particular we
suggest that Duende is now in a non-principal, short axis mode rotation state
with a precessional period equal to P1 and oscillation about the symmetry
axis at a rate equal to P2. Temporal and signal-to-noise limitations inherent
to the pre-flyby photometric dataset make it difficult to definitively diag-
nose whether these periods represent a change imparted due to gravitational
torques during the flyby. However, based on multiple analysis techniques and
a number of plausibility arguments, we suggest that Duende experienced a
rotational change during the planetary encounter with an increase in its pre-
cessional rotation period. Our preferred interpretation of the available data
is that the precession rate increased from 8.4 hours prior to the flyby to 8.7
hours afterwards. A companion paper by Benson et al. (2019) provides a more
detailed dynamical analysis of this event and compares the data to synthetic
lightcurves computed from a simple shape model of Duende. The interpreta-
tion and results presented in these two works are consistent with one another.
The ultimate outcome of this campaign suggests that the analytic tools we
4
employed are sufficient to extract detailed information about solid-body ro-
tation states given data of high enough quality and temporal sampling. As
current and future discovery surveys find more near-Earth asteroids, the op-
portunities to monitor for physical changes during planetary encounters will
increase.
Keywords: Asteroids, dynamics ; Asteroids, rotation ; Asteroids, composition
; Near-Earth objects
5
1
Introduction
It has long been understood that the physical properties of small Solar Sys-
tem bodies can be altered when passing within a few radii of a massive planet
(Roche, 1849). Numerous theoretical and observational studies suggest that
gravitationally-induced physical modification of near-Earth objects (NEOs)
can occur during close encounters with the terrestrial planets. The presence
of doublet craters and crater chains on the Earth and Venus can be explained
by tidal disruption of rubble pile progenitors during close encounters with the
Earth (Bottke & Melosh, 1996; Richardson et al., 1998). Simulations of rubble
pile asteroids suggest that close encounters approaching the Earth's Roche
limit (∼ 3 − 5 Earth radii depending on body density, 1 Earth radius = 6371
km) can result in the redistribution or stripping of surface material, the ini-
tiation of landslides, and/or overall changes in body morphology (Richardson
et al., 1998; Yu et al., 2014). The redistribution of surface material can obser-
vationally manifest as a change in spectral properties as fresh, unweathered,
sub-surface material is excavated (Nesvorny et al., 2005; Binzel et al., 2010).
Such close encounters can also increase or decrease spin rate depending on
the details of the encounter, e.g. body morphology, encounter distance, body
orientation during encounter, or body trajectory (Scheeres et al., 2000, 2004,
2005). The magnitude of these effects scales in various (and sometimes un-
known) ways as a function of encounter distance. For the specific case of the
asteroid 4179 Toutatis a small (< 1%) change in its rotational angular mo-
mentum was detected following an Earth-encounter at a distance of four times
the Earth moon separation or a distance of over 200 Earth radii (Takahashi
et al., 2013). Despite this foundation of theoretical work that describes the
role of planetary encounters in altering the physical properties of asteroids,
6
these effects have never been observed in real-time for encounters between an
asteroid and a terrestrial planet.
We will present here a campaign focused on the near-Earth asteroid 367943
Duende (2012 DA14) to probe for physical changes during its near-Earth flyby
in early 2013. As we will demonstrate, Duende is now in a non-principal axis
rotation state. The details of how an asteroid enters into non-principal axis
rotation are not well understood (Pravec et al., 2005). For near-Earth ob-
jects, tidal torques experienced during planetary encounters can induce such
excited rotation states (Scheeres et al., 2000, 2004). The exact encounter dis-
tance at which non-principal axis rotation can be induced depends sensitively
on the orientation of the asteroid relative to the planet at the time of the
flyby and the initial rotation rate of the asteroid. Percent level changes in spin
rate can be induced for encounters at distances of many tens of Earth radii
(Scheeres et al., 2000). Once in a non-principal axis state, energy is dissipated
through internal stress causing a damping of the excited rotation and evolu-
tion into a state of constant rotation about the axis of maximum moment of
inertia, i.e. principal axis rotation. The characteristic time scale for damping
of non-principal axis rotation is dependent on internal properties like rigidity
and efficiency of energy dissipation, rotation rate, and body size (Burns &
Safronov, 1973; Harris, 1994). For the population of known tumblers (Warner
et al., 2009) smaller than 100 meters (relevant to NEOs and Duende), co-
hesionless damping timescales range from greater than the age of the Solar
System to under 1 Myr (Sanchez & Scheeres, 2014). Open questions remain
regarding how internal structure, e.g. cohesionless rubble piles vs. coherent
bodies, affects these damping timescales (Sanchez & Scheeres, 2014).
Asteroid discovery surveys are currently yielding more than 100 new NEOs
7
every month and are increasingly finding smaller bodies as completeness is
reached for large objects (Galache et al., 2015). The shift towards increasingly
smaller NEOs means that these objects must be at smaller geocentric distances
to reach survey detection limits. Current surveys are finding several objects
every month that pass within at least 1 lunar distance of the Earth (1 LD =
60 Earth radii = 0.0026 AU). Typically objects that undergo these near-Earth
encounters are discovered only days before closest approach, requiring rapid
response capabilities to conduct telescopic observations. However, this was not
the case with the discovery of Duende.
On UT 23 February 2012 the La Sagra Sky Survey discovered the approxi-
mately 40 meter near-Earth asteroid 367943 Duende. Within days of discovery
it was realized that this asteroid would experience a close encounter with the
Earth a year later on 15 February 2013 around 19:25 UT (JD 2456339.3090).
This flyby would occur at an altitude of roughly 27,700 km or 4.2 Earth radii,
inside the orbital ring of geosynchronous satellites and just outside the Earth's
Roche limit. This would be the 8th closest encounter ever recorded between
an asteroid and the Earth, and the first time such an event was known about
more than a few days in advance. The recovery of the asteroid a year after
discovery in January 2013 (Moskovitz et al., 2013) and the subsequent refine-
ment of the object's orbit enabled characterization observations during this
unusual event.
To investigate the possibility of gravitationally induced changes to Duende's
physical properties we conducted an extensive campaign of spectroscopic and
photometric observations immediately after discovery in 2012 and one year
later surrounding the close encounter. The primary focus of our observational
campaign involved monitoring the rotational state of the asteroid by measur-
8
ing photometric lightcurves. We present the result of the observations in §2.
The lightcurve photometry clearly indicates that Duende is in a non-principal
axis rotation state. In §3 we perform detailed analyses of the lightcurve data,
namely least-squares fitting and computation of Fourier power spectra, on the
data collected prior to and following the near-Earth flyby. This enables phys-
ical interpretation of the spin state of the body and a comparison of the pre
and post-flyby rotation rates (§4). A companion paper (Benson et al., 2019)
theoretically explores the solid body dynamics of Duende during the planetary
encounter and finds similar results to what are presented here. These dynam-
ical models coupled with feasibility arguments founded on our data analysis
suggests that Duende may have experienced rotational changes during the
planetary encounter in 2013. The implications of this work are discussed in
Section 5.
2 Observations and Data Reduction
Our observational campaign involved photometry and spectroscopy at visible
wavelengths. We also leverage Doppler delay radar imaging and radar speckle
tracking results (Benner et al., 2013). The visible spectra were obtained to
search for spectral changes attributable to seismically induced resurfacing that
would reveal fresh, unweathered, subsurface material (Binzel et al., 2010). The
lightcurve photometry was obtained to search for rotationally induced changes
or mass loss during the planetary encounter. The radar observations constrain
the rotation state and morphology of the asteroid.
9
Telescope
Location
Instrument
UT Date
V-mag
Phase Angle
Exposures
Solar Analog
Discovery Epoch, March 2012
Gemini North 8m
Mauna Kea, HI
GMOS
2012.03.02
19.9
Pre-Flyby, February 2013
Faulkes South 2m
Siding Spring, Australia
FLYODS
2013.02.15
14.1
Post-Flyby, February 2013
Nordic Optical Telescope 2.5m
La Palma, Canary Islands
ALFOSC
2013.02.17
15.8
60◦
109◦
84◦
12 x 500s
SA107-998
2 x 180s
SA98-978
7 x 300s
HD245
Table 1
Observational Summary of Duende Spectroscopy
2.1 Visible Spectroscopy
Three visible wavelength (∼ 0.4 − 0.9 µm) spectra are presented here, two
taken prior to the flyby and one afterwards (Table 1, Figure 1). The spectral
data were obtained with GMOS (Gemini Multi-Object Spectrograph) at the
Gemini North 8m on Mauna Kea in Hawaii, FLOYDS (Folded Low-Order
Yte-pupil Double-dispersed Spectrograph) on the Faulkes South telescope at
Siding Springs Observatory in Australia, and with ALFOSC (Alhambra Faint
Object Spectrograph and Camera) at the Nordic Optical Telescope (NOT) on
La Palma in the Canary Islands. In all cases these spectra and their errors
represent an error-weighted average of multiple individual exposures.
The GMOS instrument at Gemini-North was employed 10 days after the dis-
covery of the asteroid in 2012 following a successful request for director's
discretionary time. The instrument was configured with the R150 grating, a
2.0" slit and a GG455 blocking filter, producing a useful spectral range from
0.48 to 0.92 microns at a resolution per channel of approximately 0.35 nm.
Three sets of 4 x 500s exposures were obtained with a different grating angle
for each set. The three grating angles yielded central wavelengths of 0.69, 0.7
and 0.71 microns and minimized the effects of gaps between the three detec-
tors that populate the focal plane. The GMOS slit was aligned to the default
instrument position angle of 90◦. As the observed airmass of the targets was
between 1.24 and 1.32, differential refraction was negligible, particularly with
10
Fig. 1. Normalized visible spectra of Duende including pre-flyby data from the
discovery epoch in 2012 (Gemini-GMOS, red), pre-flyby data about 9 hours before
closest approach (Faulkes-FLOYDS, black), and post-encounter data taken two days
after the flyby (NOT-ALFOSC, blue). Taxonomic envelopes from the system of Bus
& Binzel (2002) are shown in grey in the background for Ld- and K-types, an L-type
classification is intermediate to these two. The variability between spectra is most
likely systematic in nature and not due to spectral changes of the asteroid following
the planetary encounter.
a 2" slit, and thus did not necessitate aligning the slit with the parallactic
angle. Reduction of these data employed the Gemini IRAF package. These
observations were made challenging by the faintness of the object (V∼20)
coupled with its moderate non-sidereal rates of 3"/min. The individual re-
duced spectra showed significant variability in slope at the ∼15% level. The
combination of all 12 GMOS spectra in Figure 1 shows error bars that solely
account for the average measured signal-to-noise per pixel and do not repre-
sent this slope variability. The cause of this significant slope variability is not
clear. Weather conditions were generally favorable: Gemini metrics of image
quality and cloud cover were in the 70-80th percentile and 50-70th percentile
11
0.40.50.60.70.80.9Wavelength (µm)0.20.40.60.81.01.21.41.6Normalized ReflectanceLdKPost Flyby - NOTPre Flyby - GeminiPre Flyby - Faulkesrespectively. The object appeared well aligned on the slit in the acquisition
frames, though no additional images of the slit were taken to confirm align-
ment once the 12 x 500s spectral sequence began. The measured variability
was largely tied to the last 4 exposures with the grating in the 0.69 micron
setting. These exposures were taken at the lowest airmass. There were no clear
indications in these spectral images of background field star contamination,
however inspection of finder charts do indicate possible contamination from
at least one faint field star during these last few exposures. It is possible this
could be the cause of the slope variability, but that is difficult to confirm.
A second pre-flyby spectrum was obtained using the FLOYDS instrument on
the Faulkes-South telescope on the day of the flyby. FLOYDS has a fixed
configuration that produces spectra from approximately 0.4 to 0.9 microns.
Spectra are captured in two orders, a red order covering 0.6 to 0.9 microns at a
per-pixel resolution of 0.35 nm and a blue order covering 0.4 to 0.6 microns at
a per-pixel resolution of 0.17 nm. A 2" slit was used for these observations. We
re-binned and spliced the red and blue orders to increase signal-to-noise, pro-
ducing a final spectrum at a resolution of 0.01 microns (Figure 1). The high
non-sidereal motion of the asteroid during these observations (>80 "/min)
made acquisition sufficiently challenging that only two spectra were success-
fully captured. In one case the object clearly drifted off of the slit during
the exposure as indicated by significantly reduced counts. As such the slope
of the final spectrum may have been affected. The slit was aligned with the
parallactic angle for these observations.
The ALFOSC instrument at the NOT was used 2 days after the flyby on 17
February 2013. The instrument was configured with the #11 grism and a 1.8"
slit, producing a useful spectral range from 0.40 to 0.87 µm at a resolution
12
per channel of approximately 0.47 nm. Reduction of these data employed
the IRAF specred package (Tody, 1986). The observing conditions on this
night were not good with high humidity, clouds, and poor seeing. Individual
spectra showed variability in slope at the ∼10% level. The combination of
all 7 ALFOSC spectra in Figure 1 shows error bars that solely account for
the average measured signal-to-noise per pixel and do not represent this slope
variability. The ALFOSC slit was aligned with the parallactic angle.
The spectra are best fit with an Ld-, K-, or L-type classification in the Bus
& Binzel (2002) system. These taxonomic assignments are consistent with the
findings of de Leon et al. (2013) and Takahashi et al. (2014). The composition
of these taxonomic types is not well understood, though suggested connections
to carbonaceous chondrite meteorites have been made (Bell, 1988). This am-
biguous composition complicates our search for spectral changes that might be
attributed to exposure of fresh, sub-surface material. Carbonaceous chondrites
are not generally associated with traditional lunar-style space weathering that
causes an increase in spectral slope and a decrease in absorption band depth
(Hapke, 2001). In fact, some studies find that carbonaceous meteorites get
spectrally bluer during simulated space weathering processes (Lantz et al.,
2013). Thus, the composition of Duende may spectroscopically weather in a
manner that is not currently well understood. Regardless of this ambiguity,
we attribute the observed spectral slope variation (Figure 1) to observational
systematics -- e.g. improper alignment of the asteroid on the slit, airmass
differences between the asteroid and calibration solar analog star -- and not
physical changes on the asteroid's surface. Though there are differences in the
slopes of the final spectra presented here (Figure 1), the variation amongst
individual exposures from a single instrument generally span an appreciable
fraction of this slope range. Furthermore, an instrumental or observational
13
cause of slope variation is supported by the pronounced difference in spectral
slope between the two pre-flyby spectra, taken before any potential surface
perturbation could have occured. Even though these observations were taken
across a range of phase angles (Table 1), there is no clear signature of slope
variability that could be attributed to phase reddening (Sanchez et al., 2012).
Therefore we conclude that this slope variability is systematic in nature and
not due to spectral changes intrinsic to the asteroid. This suggests that the
surface was not significantly perturbed during the flyby and thus did not reveal
unweathered subsurface material, or the composition of Duende is insensitive
to space weathering effects at visible wavelengths, or any spectral changes were
below the sensitivity of our measurements. This result is fully consistent with
the non-detection of near-infrared color changes during the flyby (Takahashi
et al., 2014).
2.2 Photometry
The primary focus of our observational campaign involved measuring rota-
tional lightcurves via broadband photometry. These observations were con-
ducted in March 2012 following the discovery of Duende and in February
2013 on either side of the close encounter. In total these observations incorpo-
rated data from 15 different observatories. Examples of images from a number
of these observatories are shown in Figure 2. The orbital geometry of the
flyby was such that the asteroid approached from a large negative declina-
tion of approximately -75◦ and receded at a large positive declination around
+80◦. Thus, pre-flyby characterization required observations conducted from
the southern hemisphere, while post-flyby observations were conducted from
the North (Figure 3). These observations spanned effectively the full extent
14
of the observable sky: more than 300◦ in right ascension and more than 150◦
in declination. In the hours surrounding the flyby, a rapidly changing view-
ing geometry and the lack of a detailed shape model would have made it
difficult to link the body's physical rotation state to measured lightcurves.
Thus, our strategy was to compare rotation state before and after the flyby
to look for evidence of spin changes. Details of the lightcurve observations are
summarized in Table 2 and each observatory/instrument combination is de-
scribed below. In general reduction of our photometric data followed standard
aperture photometry techniques in IRAF following the methods of Moskovitz
(2012), though several exceptions are noted. All data were transformed to
V-band magnitudes (Dandy et al., 2003) based on reference magnitudes of
on-chip field stars. The temporal overlap of lightcurve segments from multiple
observatories (Table 2) provides confirmation that our magnitude calibration
techniques are reasonable and accurate to < 0.1 magnitudes. For analysis pur-
poses all data have been corrected for geocentric range and phase angle using
an ephemeris generated with the JPL Horizons system so that we only work
with differential magnitudes relative to the ephemeris predictions.
15
Fig. 2. Example images from several of the instruments used in the photometry
campaign. Each field is approximately 3' wide and all are orientated with North up,
East left as indicated by the compass in the upper right. Duende is circled in each
frame, except for the image from the Perth 0.35m, where the polygonal photometric
aperture is shown.
16
Fig. 3. Geocentric coordinates surrounding the dates of closest approach. Each point
represents one of our lightcurve observations, both pre (light grey) and post (dark
grey) encounter are included. The start and end times of the pre and post-flyby
sequences are shown.
17
060120180240300360Right Ascension (deg)7550250255075Declination (deg)Pre2013-Feb-09 09:472013-Feb-15 16:15Post2013-Feb-16 00:002013-Feb-19 19:48i
e
m
T
n
o
i
t
a
v
r
e
s
b
O
)
0
0
0
5
5
4
2
-
D
J
(
r
e
t
l
i
F
e
t
a
l
P
e
l
a
c
S
D
C
C
e
z
i
S
t
n
e
m
u
r
t
s
n
I
n
o
i
t
a
c
o
L
e
p
o
c
s
e
l
e
T
2
1
0
2
h
c
r
a
M
,
h
c
o
p
E
y
r
e
v
o
c
s
i
D
2
9
6
5
1
.
9
8
9
-
7
4
3
0
9
.
8
8
9
I
"
7
3
1
.
0
k
4
x
k
4
I
C
T
P
O
I
H
,
a
e
K
a
n
u
a
M
m
2
.
2
i
i
a
w
a
H
f
o
y
t
i
s
r
e
v
i
n
U
6
2
4
0
0
.
5
3
3
1
1
6
8
7
5
.
4
3
3
1
1
4
6
5
6
.
4
3
3
1
9
6
8
4
5
.
7
3
3
1
8
6
6
2
4
.
8
3
3
1
0
0
7
7
9
.
8
3
3
1
7
3
7
7
1
.
9
3
3
1
8
5
7
5
6
.
9
3
3
1
6
4
8
4
7
.
9
3
3
1
8
9
7
5
0
.
0
4
3
1
1
3
4
1
5
.
0
4
3
1
4
7
5
3
7
.
1
4
3
1
1
0
4
5
0
.
2
4
3
1
-
-
-
-
-
-
-
-
-
-
-
-
-
1
1
2
5
0
.
9
3
3
1
n
e
p
O
"
6
6
.
0
k
5
.
1
x
k
2
P
O
C
R
a
i
l
a
r
t
s
u
A
,
h
t
r
e
P
m
5
3
.
0
y
r
o
t
a
v
r
e
s
b
O
h
t
r
e
P
3
3
0
0
5
.
9
3
3
1
9
7
8
9
5
.
9
3
3
1
n
e
p
O
R
9
5
8
7
9
.
9
3
3
1
'
g
S
S
D
S
3
4
2
8
3
.
0
4
3
1
7
1
1
3
6
.
0
4
3
1
R
R
"
1
.
1
k
5
.
1
x
k
2
D
C
C
G
I
B
S
l
e
a
r
s
I
,
v
i
v
A
l
e
T
m
6
4
.
0
y
r
o
t
a
v
r
e
s
b
O
e
s
i
W
"
8
3
.
0
k
4
x
k
4
k
4
T
T
A
V
a
n
o
z
i
r
A
,
m
a
h
a
r
G
.
t
M
m
8
.
1
T
T
A
V
"
3
.
0
"
2
.
2
k
2
x
k
2
B
K
2
T
a
n
o
z
i
r
A
,
k
a
e
P
t
t
i
K
m
1
.
2
k
a
e
P
t
t
i
K
k
1
x
k
1
D
C
C
e
e
g
o
p
A
d
n
a
l
g
n
E
,
e
r
i
h
s
k
r
e
B
t
s
e
W
y
r
o
t
a
v
r
e
s
b
O
d
r
o
ff
e
h
S
t
a
e
r
G
"
3
5
.
0
k
2
x
k
2
r
e
g
a
m
I
k
2
o
c
i
x
e
M
w
e
N
,
o
r
r
o
c
o
S
m
4
.
2
i
e
g
d
R
a
n
e
l
a
d
g
a
M
3
1
0
2
y
r
a
u
r
b
e
F
,
y
b
y
l
F
-
t
s
o
P
0
2
4
8
7
.
1
4
3
1
'
r
S
S
D
S
"
1
1
.
0
k
2
1
x
k
2
1
I
D
O
p
a
n
o
z
i
r
A
,
k
a
e
P
t
t
i
K
m
5
.
3
N
Y
W
I
2
e
l
b
a
T
0
9
0
3
.
9
3
3
6
5
4
2
D
J
t
a
r
e
t
n
u
o
c
n
E
e
s
o
l
C
e
h
t
g
n
i
d
n
u
o
r
r
u
S
y
r
t
e
m
o
t
o
h
P
e
d
n
e
u
D
f
o
y
r
a
m
m
u
S
l
a
n
o
i
t
a
v
r
e
s
b
O
8
8
6
9
8
.
2
3
3
1
n
e
p
O
3
2
6
4
5
.
4
3
3
1
0
0
8
7
-
0
0
8
4
B
W
"
3
.
1
"
2
.
0
k
1
x
k
1
D
C
C
e
e
g
o
p
A
d
n
a
l
a
e
Z
w
e
N
,
o
p
a
k
e
T
e
k
a
L
m
1
y
r
o
t
a
v
r
e
s
b
O
n
h
o
J
.
t
M
k
8
x
k
8
S
C
A
M
I
e
l
i
h
C
,
s
a
n
a
p
m
a
C
s
a
L
m
5
.
6
e
d
a
a
B
n
a
l
l
e
g
a
M
3
1
0
2
y
r
a
u
r
b
e
F
,
y
b
y
l
F
-
e
r
P
9
0
6
1
6
.
4
3
3
1
w
-
S
R
R
A
T
S
n
a
P
"
2
3
2
.
0
k
4
x
k
4
D
C
C
G
I
B
S
e
l
i
h
C
,
o
l
o
l
o
T
o
r
r
e
C
6
6
0
2
5
.
7
3
3
1
'
r
S
S
D
S
"
9
5
2
.
0
k
2
x
k
2
D
C
C
k
2
e
T
I
S
e
l
i
h
C
,
s
a
n
a
p
m
a
C
s
a
L
0
2
2
0
4
.
8
3
3
1
8
8
7
0
9
.
8
3
3
1
J
R
V
"
6
4
.
0
k
1
x
k
1
I
S
U
R
I
S
a
c
i
r
f
A
h
t
u
o
S
,
O
A
A
S
"
3
.
0
k
4
x
k
4
l
a
r
t
c
e
p
S
a
i
l
a
r
t
s
u
A
,
g
n
i
r
p
S
i
g
n
d
S
i
m
2
h
t
u
o
S
s
e
k
l
u
a
F
m
1
T
G
O
C
L
m
5
.
2
t
n
o
P
u
D
m
4
.
1
F
S
R
I
18
University of Hawaii 2.2m: We employed the Orthogonal Parallel Transfer
Imaging Camera (OPTIC) on the University of Hawaii 2.2 meter telescope to
obtain Cousins I-band photometry of Duende within a week of its discovery.
OPTIC consists of two 2k x 4k Lincoln Lab detectors covering a field of view
roughly 9.3' x 9.3'. The telescope was tracked at half the non-sidereal rate
of the asteroid to minimize elongation of the field stars and asteroid in each
exposure such that circular aperture photometry could be performed. We used
an exposure time of 120 seconds. Reduction of these data employed standard
IRAF routines.
Mt John Observatory 1m: We employed an Apogee F6 1k x 1k KAF1001E
CCD with 24 µm pixels and a plate scale of 0.64". The detector was binned
2 x 2 and exposure times were set to 60 seconds. The telescope was guided at
sidereal rates allowing the asteroid to move through the fixed 11' x 11' field
of view. Reduction of these data employed standard IRAF tools.
Magellan Baade 6.5m: Observations were conducted with IMACS (Dressler et
al., 2011) at the Magellan Baade 6.5 m telescope at Las Campanas Observatory
in Chile. IMACS was operated using the f/2 camera, which has a 27.50 field-of-
view covered by a mosaic of eight 2k x 4k CCDs with plate scales of 0.2"/pixel.
Exposure times of 7-10 seconds and a broadband filter with spectral response
from 0.48 - 0.78 µm were used. Reduction of these data employed standard
IRAF tools.
Las Cumbres Observatory Global Telescope Network (LCOGT) 1m: We em-
ployed a 4k x 4k SBIG CCD with a 0.232" plate scale covering a 15.8' field
of view and an exposure time of 180 seconds. The frames were binned 2 x 2.
Data from this observatory and the Faulkes South 2m (described below) were
pre-processed through the LCOGT pipeline (Brown et al., 2013) to perform
19
bias and dark-subtraction, flat fielding, and to determine an astrometric solu-
tion against the USNO-B1.0 (Monet et al., 2003) for the 2m and the UCAC-3
(Zacharias et al., 2010) catalog for the 1m. The 1m photometry were mea-
sured using SExtractor (Bertin & Arnouts, 1996) to fit elliptical apertures to
the trailed, elongated asteroid images.
DuPont 2.5m: Observations were performed at the DuPont 2.5 m telescope
at Las Campanas Observatory with a SITe2k CCD. The SITe2k at DuPont is
a 2k x 2k CCD with an 8.85' field-of-view and a plate scale of 0.259"/pixel.
Exposures of 60 seconds were obtained. Reduction of these data employed
standard IRAF tools.
Infrared Survey Facility (IRSF) 1.4m: Observations were obtained with the
Simultaneous three-color Infrared Imager for Unbiased Surveys (SIRIUS) at
the IRSF located in Sutherland, South Africa. SIRIUS consists of three 1k x
1k Hawaii HgCdTe arrays (for simultaneous J, H and K band imaging) with
plate scales of 0.453"/pixel, each covering a 7.7' field of view.
Faulkes South 2m Telescope: We employed the Spectral instrument which con-
tains a 4k x 4k camera for imaging. The plate scale of Spectral is 0.152" and
the field of view is 10.5'. Frames were binned 2 x 2. Exposure times ranged
from 180 seconds on UT 9 February 2013 to 7 seconds on 15 February 2013.
The frames were processed in Astrometrica using the UCAC-3 catalog for
astrometric reference and then positions and magnitudes were determined.
Perth Observatory R-COP 0.35m: Observations were conducted on the day
of the flyby with the 0.35m R-COP telescope at Perth Observatory using
a 2k x 1.5k SBIG CCD with 0.66" pixels and a 24.7' x 16.5' field of view.
All images were obtained unfiltered. Exposure times ranged from 10 seconds
20
at the onset (when the object was moving slowest) to 1 second about three
hours prior to closest approach when the object was moving at non-sidereal
rates in excess of 500 "/minute. These were the only data for which we did
not employ standard circular or elliptical aperture photometry techniques.
The high non-sidereal rate of the asteroid in these images resulted in trailing.
As such we employed polygonal aperture photometry with the IRAF polyphot
routine. Field stars in these images were examined to determine the seeing; the
width of the polygonal aperture along the short (untrailed) axis of the asteroid
was interactively defined to equal approximately twice the local seeing. The
polygonal apertures were defined with six points, three at each end of the
trailed asteroid. The background annulus was defined with a 60-pixel inner
radius and a 30-pixel width.
Wise Observatory 0.46m: The 0.46m at Wise Observatory in Israel (Brosch
et al., 2008) was employed about 4.5 hours after closest approach. The instru-
ment was a 2k x 1.5k SBIG CCD with 1.1" pixels and a 40.5' x 27.3' field
of view. Images were obtained unfiltered. The telescope pointing was offset
between sets of exposures to keep up with the fast motion of the asteroid.
This was conducted automatically throughout the night of 2013 February 15
by programming the telescope to follow ephemeris coordinates from the Minor
Planet Center (MPC). IRAF's phot function with an aperture of 4 pixels was
used for measuring the photometry. The differential magnitude of the asteroid
was calibrated based on comparison to hundreds of on-chip field stars. The
brightness of these stars remained constant to within ±0.02 mag. See Pol-
ishook & Brosch (2009) for details of the reduction algorithm. Nine different
fields-of-view were used to track the asteroid over the course of about 4.5
hours. This resulted in nine lightcurve segments that were stitched together
to form a continuous lightcurve.
21
Vatican Advanced Technology Telescope (VATT) 1.8m: The VATT 1.8m was
operated with the VATT 4k STA CCD. The CCD covers a 12.5' square field
of view at 0.188 "/pixel. Exposure times ranged from 2 to 5 seconds and the
telescope was tracked at the non-sidereal rates of the asteroid. Trailing of the
field stars was minor thus circular aperture photometry was possible.
Kitt Peak 2.1m: The T2KB Tektronix 2k x 2k CCD was employed at the
Kitt Peak 2.1m. The plate scale of this instrument is 0.3"/pixel and the un-
vignetted field of view of 10.2' x 9.4'. We obtained 5 second exposures and
guided the telescope at sidereal rates allowing the asteroid to move through
fixed star fields. Reductions and photometry followed standard IRAF proto-
cols.
Great Shefford Observatory 0.4m: The 0.4m Schmidt-Cassegain telescope at
Great Shefford Observatory in Berkshire, England was used with a 1k x 1k
E2V CCD. The plate scale of the CCD is 1.1 "/pixel and produces a vignetted
18' circular field of view. Exposure times of 15 to 30 seconds were used. Re-
duction of the photometry was performed using Astrometrica referenced to
the PPMXL (Roeser et al., 2010) catalog.
Magdalena Ridge Observatory (MRO) 2.4m: The MRO 2.4m was employed
with a 2k x 2k CCD with a 0.53"/pixel plate scale that images an 18' square
field of view. Observations were made tracking on Duende while periodically
shifting to a nearby comparison field. Seeing was typically 1.0-1.2" on UT 17
February 2013 and 1.5-1.9" on 18 February 2013. Data were measured using
IRAF's apphot package with 12 pixel and 16 pixel apertures on 17 February
and 18 February respectively (large apertures allowed for seeing variations).
The comparison fields were calibrated using Landolt standards and differential
photometry was performed using a custom program assuming a V-R of 0.47 for
22
the target, roughly consistent with Duende's L-type taxonomic classification
(Dandy et al., 2003).
WIYN 3.5m: The WIYN 3.5m was employed with the partial One Degree
Imager (pODI; Harbeck et al., 2014) which is a mosaic of multiple 4k x 4k
STA orthogonal transfer CCDs. These chips have a plate scale of 0.11"/pixel
and each covers an 8' square field of view. We employed the central 9 CCDs
to produce an overall field 24' x 24'. The wide field-of-view of pODI allowed
us to use a single pointing for the entire night of UT 17 February 2013. This
enabled use of a single set of reference field stars for measuring the differential
magnitude of the asteroid. Many images were excluded from analysis because
of the proximity of the asteroid to pODI's numerous chip and cell boundaries.
Reduction of the data employed a custom set of python routines (courtesy R.
Kotulla). Photometry was performed using IRAF's apphot package with an
aperture of 15 pixels, and a background annulus with radius 30 pixels and
width of 25 pixels.
In addition to the photometry obtained as part of our campaign, we used pub-
licly available data from Gary (2013) and a few select measurements submitted
to the MPC. We only included sets of MPC data for which there were 3 or
more observations within a single night and when the data did not show large
(>0.2 magnitude) random fluctuations (Table 3). These criteria removed the
vast majority of the more than 1000 individual MPC observations of Duende,
adding only 11 data points to our pre-flyby lightcurve and 161 to our post-
flyby curve.
Our post-flyby images showed no evidence for mass loss or mass shedding.
No difference was seen in the width of Duende's point-spread function (PSF)
compared to that of field stars in the same images, though a detailed search
23
Observatory
MPC Code
UT Date
Pre-Flyby
2013.02.14
2013.02.15
2013.02.15
Post-Flyby
2013.02.16
2013.02.16
2013.02.16
Cerro Tololo Observatory
Arcadia
Murrumbateman
Kurihara
Pulkova
Montevenere Observatory
807
E23
E07
D95
084
C91
J38
300
958
168
H06
B67
2013.02.16, 2013.02.19 Observatorio La Vara, Valdes
2013.02.16
2013.02.17
2013.02.17
2013.02.18
2013.02.18
Bisei Spaceguard Center - BATTeRS
Observatoire de Dax
Kourovskaya
iTelescope Observatory, Mayhill
Sternwarte Mirasteilas, Falera
Table 3
Summary of MPC Data Used in lightcurve Fits
for faint extended features (e.g. Sonnett et al., 2011) was not conducted. The
lack of clear signatures for mass loss is consistent with post-flyby radar data
that were sensitive enough to resolve, but did not detect, the escape of sub-
meter-scale fragments from the surface (Benner et al., 2013).
2.3 Radar Constraints
Post-flyby radar imaging and speckle tracking (Benner et al., 2013) provide
important constraints, that in conjunction with the lightcurve analysis, allow
us to better understand the post-flyby shape and spin state. The delay-Doppler
images are consistent with a multi-hour rotation state around 8 hours, but
they do not provide a sufficiently long interval to uniquely constrain any non-
principal axis rotation. As we show in the following sections, the lightcurve
photometry clearly shows that Duende is in non-principal axis rotation. The
radar imaging suggests that the body is in a relatively flat spin, which is most
consistent with non-principal short-axis mode (SAM) rotation as opposed to a
long-axis mode (LAM) precession state (Kaasalainen, 2001), of which asteroid
24
4179 Toutatis is a well known example (Ostro et al., 1995; Hudson & Ostro,
1995). Other NEOs with SAM rotation states include 99942 Apophis (Pravec
et al., 2014) and (214869) 2007 PA8 (Brozovi´c et al., 2017). Radar speckle
tracking from the Very Large Array (VLA) constrain the amplitude of non-
principal axis wobble (i.e. the long axis nutation angle) to < 30◦, which is
also consistent with a SAM state. A more detailed discussion of the body
dynamics is presented in Benson et al. (2019). The highest resolution images
from the Goldstone array in California achieved 3.75m per pixel resolution
and determined that the asteroid is highly elongated (approximately 40 × 20
meters), though no surface features were resolved.
3 Lightcurve Analysis
As discussed above, lightcurve photometry was obtained during three win-
dows: in 2012 during the discovery apparition, in 2013 before the time of closest
approach (19:25 UT 15 February 2013), and in 2013 following the flyby. The
radar data were collected in the days following closest approach. We focused
our observations and analysis on data taken more than 3 hours before and 3
hours after the time of closest approach. This restriction was an intentional
part of our experimental design and had a number of important benefits. First,
avoiding the time of closest encounter ensured that the non-sidereal rates of
the asteroid were an order of magnitude lower than at their peak. This made
the observations and reductions simpler because significant trailing of the as-
teroid was avoided and a relatively small number of fields could be used for
differential photometry against background field stars. Second, this strategy
ensured that the viewing geometry did not change significantly during the
windows of observation. This can be quantified in terms of the solar phase an-
25
gle, which changed by < 25◦ during the inbound observing window and < 20◦
outbound. Third, this strategy was based on models suggesting that the ma-
jority of gravitationally-induced changes in rotation state occur within a few
hours of closest approach (Scheeres et al., 2000). Thus, our observing windows
served to isolate the pre- and post-flyby rotation states of the asteroid, inde-
pendent of any potential changes. Fourth, conducting our observations in these
discrete windows meant that the difference in sidereal versus synodic rotation
periods caused insignificant phase shifts of <10 seconds. Here we describe the
resulting lightcurves and the fitting of those data to derive constraints on the
rotational state of the asteroid. Specifically, we perform both a minimum-RMS
analysis and a fast Fourier transform (FFT) analysis on each of the data sets
to facilitate comparison of the pre and post-flyby rotation states.
3.1 Post-Flyby
The post-flyby lightcurve starts about 4.5 hours after closest approach, pro-
vides the strongest constraints on the rotational properties of Duende, and
serves as a baseline for our analysis (Figure 4). The more than 3000 individ-
ual data points are densely sampled with several segments overlapping in time,
which provides a nice check on our magnitude calibrations. The non-repeating
morphology of the post-flyby lightcurve is a clear indication of non-principal
axis rotation (informally known as tumbling) and can be modeled with a two
dimensional (dual period) Fourier series. This model follows the formalism of
Pravec et al. (2005) and was applied to the combination of our observations
and the selected data from the MPC. This procedure involved determining an
error weighted least squares fit to the data with the IDL mpfit package using a
second order 2D Fourier series (Equation 6 in Pravec et al., 2005) as the input
26
function. This function includes 27 free parameters: an overall magnitude off-
set relative to zero, two distinct periods labeled P1 and P2, and 24 coefficients
accounting for the amplitudes of each Fourier term. With the exception of
the initial conditions for the two periods, all parameters in our model were
initially set to zero. We also restricted the individual Fourier coefficients to
< 0.5 magnitude to limit net amplitude in the final functional fit. As noted
in Pravec et al. (2005) and described in greater detail below the periods P1
and P2 can not always be assigned to specific rotational modes of the body
due to issues of aliasing. We also tested 1st, 3rd and 4th order Fourier series
which produced equivalent results. We settled on a second order function as a
compromise between minimizing the number of free parameters in the model
and providing a reasonable fit to the data. Small vertical offsets within the
signal-to-noise of each individual data set were manually applied to minimize
the chi-squared residual of the model fits. Ultimately this approach produced
strong constraints on the fitted periods, even though the fine-scale structure
in the peaks and troughs of the lightcurve were not always perfectly fit.
The criteria for using MPC data (Table 3) in this analysis was as follows:
we first determined a best-fit lightcurve to our observations only (filled dots
in Figure 4) while excluding the MPC data. The selected MPC data were
then added in with observatory dependent magnitude offsets to account for
the lack of precise photometric calibration that can be characteristic of MPC
astrometric observations. These offsets were applied to bring the MPC data
into agreement with the initial best-fit lightcurve model. We then refit the
entire data set (ours + MPC) with errors of ±0.1 magnitude assigned to the
MPC points. The MPC data did not dramatically change the model results
but did provide some temporal coverage at times where we were not able to
obtain our own observations.
27
Fig. 4. Post-flyby rotational lightcurve of Duende clearly demonstrating a non-prin-
cipal axis rotation state. The filled circles represent our observations; open diamonds
represent data reported to the Minor Planet Center. The data are well fit with a
dual period, tumbling asteroid model (grey curve), in this example with periods
P1 = 8.71 ± 0.03 and P2 = 23.7 ± 0.2 hours.
To identify viable period solutions we performed a grid search across periods
P1 and P2 from 1 to 40 hours in step sizes of 0.1 hours. At each grid point, the
periods were held fixed while the Fourier coefficients were allowed to float, and
an RMS residual (normalized by the number of data points) was computed
for each period pair. A contour plot of those RMS residuals is shown in Figure
5. This analysis reveals the three lowest RMS solutions tabulated in Table
4. The RMS residuals for these three solutions are sufficiently similar, i.e. the
differences are much less than the noise in our data, that we do not find a strong
preference for one period pair over the others. No other solutions between
periods of 1-40 hours were found with RMS residuals less than 0.1 magnitude.
Based on plausibility arguments presented here (§4) and the complementary
dynamical analysis of Benson et al. (2019), we identify the period pair of 8.71
and 23.7 hours as the preferred interpretation of Duende's post-flyby spin
state. The fit to the data shown in Figure 4 corresponds to these periods
P1 = 8.71 ± 0.03 and P2 = 23.7 ± 0.2 hours, with errors bars determined
28
01234Days relative to 00:00 UT, 16 February 20131.00.50.0-0.5-1.0Relative MagnitudeFig. 5. Contours of RMS residuals of 2D Fourier series fits for a range of post-flyby
P1 -- P2 period pairs. The RMS values are symmetric about the diagonal because the
functional form of the Fourier series model is symmetric with regards to the two
periods. The three lowest RMS solutions are marked with red dots, with the point
for our preferred solution at P1 = 8.71 hr and P2 = 23.7 hr shown larger than the
other two. These are the only solutions with RMS values below 0.1 mag.
based on the range of periods in which RMS residuals remained below 0.1
magnitude.
P1, P2
(hr)
RMS
(mag)
fa
fb
fc
fd
(1.01 day−1)
(3.78 day−1)
(5.46 day−1)
(7.61 day−1)
6.36, 8.73
0.067
8.71, 23.7
0.079
6.37, 23.7
0.089
-
1/P2
1/P2
1/P1
1/P1 + 1/P2
2/P2
2/P1
2/P1
2/P1 + 2/P2
1/P1
2/P1 - 2/P2
2/P1
Table 4
Top three period pair solutions to the post-flyby data. The corresponding frequency
peaks in our FFT analysis are assigned to first and second order linear combinations
of these periods. The preferred solution of 8.71 and 23.7 hours is highlighted.
Previous analyses of Duende's post-flyby rotation state did not span a long
enough temporal baseline to resolve complex rotation and/or were measured
29
110P1 (hr)110P2 (hr)0.050.100.150.200.250.30RMS (mag)110 110 so near to close encounter that viewing aspect was highly variable which likely
influenced the derived periods (Gary, 2013; de Leon et al., 2013). However, in
both of these previous analyses a single period fit suggested a period around
9 hours, roughly consistent with the multi-hour periods in our more detailed
analysis.
To complement our least-squares fitting of the post-flyby lightcurve, we also
perform an FFT periodogram analysis of the data (Figure 6). This analy-
sis involved computing a discrete Fourier transform of the lightcurve data
to produce a "dirty" power spectrum, which includes aliases and other un-
wanted periodicities that, for example, may reflect the temporal sampling of
the lightcurve based on observational cadences. This power spectrum was then
Fig. 6. FFT power spectra of post-flyby lightcurve. Both a clean (bottom) and dirty
(top) version of the FFT are presented, see text for more details. The four highest
peaks in the cleaned FFT are labeled in order of strength fa−d and can in most
cases be directly mapped to the periods and linear combinations of those periods
identified in our RMS fits to the data.
30
024681012cycles / day0.000.010.020.030.040.05Powerdirty024681012cycles / day0.000.010.020.03Powercleanfafbfcfd"cleaned" using the WindowCLEAN algorithm, which numerically applies a
deconvolution kernel to remove aliases in the power spectrum (Roberts et al.,
1987; Belton & Gandhi, 1988; Mueller et al., 2002). The resulting clean spec-
trum in Figure 6 clearly displays four prominent peaks that in most cases
correlate well with the derived P1 and P2 from the RMS analysis. In the Win-
dowCLEAN frequency space, peaks manifest at integer multiples or low order
linear combinations of the periods present in the data. We assign the primary
peaks in this FFT, labelled fa−d, to specific periods and linear combinations of
periods in Table 4. For example, the two strongest peaks fc = 5.46 day−1 and
fd = 7.61 day−1 closely correspond to the two periods in the minimum RMS
solution, where 2/P2 = 5.5 day−1 and 2/P1 = 7.6 day−1. In this case the peak
at fa = 3.8 day−1 is likely an alias at approximately 1/P1. The presence of
peaks at these specific aliases and linear combinations is fully consistent with
analyses of lightcurves for other non-principal axis rotators (e.g. Kryszczynska
et al., 1999; Mueller et al., 2002; Samarasinha & Mueller, 2015). In the dirty
power spectrum, a prominent peak is seen around 6.5 cycles per day, which is
then is removed by the cleaning algorithm. We interpret this as a combined
alias of the peaks at 5.46 and 7.61 cycles per day. The bi-modal morphology of
the dirty 6.5 cycle per day feature is consistent with this interpretation. Gen-
erally, major peaks in an FFT power spectrum will have symmetric aliases on
either side of the primary frequency. Post-cleaning changes in the FFT power
around 4, 6.5, and 9 cycles per day are consistent with such aliases connected
to peaks c and d.
31
3.2 Pre-Flyby and Discovery Epoch
Numerous issues related to weather and instrument failure resulted in less
extensive lightcurve coverage during the discovery and pre-flyby epochs. As
a result the lightcurve of Duende is not well constrained prior to the Earth
encounter in February 2013.
Data were acquired with the SNIFS instrument at the University of Hawaii
2.2m telescope on Mauna Kea within a week of Duende's discovery (Figure
7). These data did not span a sufficiently long baseline to resolve complex
rotation, though a single period solution suggests a period of 5.8 ± 0.3 hours.
However, a single period solution over such a short time interval for an object
Fig. 7. Pre-flyby lightcurve of Duende taken during the discovery epoch in March
of 2012. A single-period fit suggests a rotation period of approximately 5.8 hours.
However, this is not an accurate representation of an object in non-principal axis
rotation.
32
01234561818.218.418.618.81919.2Time [hours]Relative mag 2012 DA14Period=5.8±0.3 hours2 Mar 2012 UTin complex rotation is undoubtedly misleading (Pravec et al., 2005).
Pre-flyby observations in February 2013 were conducted from a number of
observatories in the southern hemisphere (Table 2). These observations inter-
mittently sampled a time span starting approximately 6.5 days before closest
approach (Figure 8). We show in this figure representative fits to the data
using the same dual period Fourier model (Pravec et al., 2005) with periods
1 = 4.5 hours and P (cid:48)
P (cid:48)
period paring of P (cid:48)
2 = 16.4 hours, our lowest RMS fit, and the preferred
2 = 24.18 hours (§4). Data from in-
dividual observatories have been iteratively offset to minimize the residual of
1 = 8.37 hours and P (cid:48)
the fit. The three data sets from Mt. John Observatory, around days -6.5, -5.5
and -4.5, are about an order of magnitude noisier than our other data sets and
thus have been temporally re-binned to an interval of ∼0.02 hours. For fitting
purposes, the errors for these time averaged points were set to the standard
deviation of the binned data.
To analyze the pre-flyby data, we employed the same 2D Fourier series function
and the same 0.1-hr period grid search as was done for the post-flyby data
(§3.1). The result of this grid search is shown in Figure 9 with the contours
shown on the same color bar. Even though this figure shows that there is no
clear single solution to the pre-flyby data, there are a number of key aspects
that provide clues about the pre-flyby spin state. First, the consistently high
RMS values along the diagonal (and at integer multiples of the diagonal) is
a clear indication that Duende was in non-principal axis rotation prior to the
encounter, i.e. two periods are needed to fit the pre-flyby data. Second, the
RMS contours are not randomly distributed, but instead show clear structure.
Of the 160,801 period combinations presented in this figure, 778 of them have
RMS residuals less than 0.1 magnitude, where the lowest at P (cid:48)
1 = 4.5 and P (cid:48)
2 =
33
Fig. 8. Pre-flyby rotational lightcurve of Duende. The time of closest approach is
labeled as ∆min (dashed line). The filled black circles represent our observations;
the open diamonds represent data reported to the Minor Planet Center. Low S/N
data (grey dots) have been combined to produce time-averaged photometric points.
The large gaps in rotational coverage make it difficult to uniquely constrain the
lightcurve in the same manner as the post-flyby data. Dual period fits to the data
are shown for the lowest RMS solution P (cid:48)
2 = 16.4 hours (dotted),
and our preferred solution of P (cid:48)
1 = 4.5 hours and P (cid:48)
1 = 8.37 hours and P (cid:48)
2 = 24.18 hours (solid grey).
16.4 hr has an RMS = 0.059. This large number of local minima do appear
in several well-defined horizontal and vertical bands (reflected symmetrically
about the diagonal) with prominent examples centered around 4.1, 6.2, 8.3,
12.5, 16.5, and 24.6 hours. The width of these bands and many of the local
minima are greater than the grid spacing of 0.1 hours, thus suggesting that
these features are a representation of real periodicity and are not dominated
by fluctuations induced by noise in the data. However, this does suggest that
the combination of noise in the data and grid spacing leads to pre-flyby period
solutions that are precise to no better than ∼ 0.1 hour.
34
-6.5-6.0-5.5-5.0-4.5Days relative to 00:00 UT, 16 February 20131.00.50.0-0.5-1.0Relative Magnitude-2.0-1.5-1.0-0.50.0Days relative to 00:00 UT, 16 February 20131.00.50.0-0.5-1.0Relative Magnitude6minFig. 9. Contours of RMS residuals of 2D Fourier series fits for a range of pre-flyby
P (cid:48)
1 -- P (cid:48)
2 period pairs. The preferred post-flyby solution of P1 = 8.71 and P2 = 23.7
hours is shown as the large red dot and falls in a region where there are no low-RMS
pre-flyby solutions. The smaller red dots are the other two post-flyby solutions
(Table 4). Though there is no single combination of periods that provide a unique
solution to the pre-flyby data, there are a large number of local minima that could be
valid. The global minimum is indicated with the white star. The white box centered
around P (cid:48)
2 = 23.7 encloses a set of low RMS solutions that represent our preferred
interpretation based on dynamical arguments (Benson et al., 2019).
In general, for the pre-flyby case, we find that a number of solutions are more-
or-less statistically indistinguishable (∼ 0.01 magnitude difference in residuals)
with RMS values approaching the noise level of our data (∼ 0.1 magnitudes).
Several of these solutions are consistent with the peaks in a WindowClean
Fourier analysis presented below. In addition, a number of the period pairs
identified in this RMS analysis are consistent with the allowable spin states
expected from a detailed analysis of Duende's solid body dynamics (Benson et
al., 2019). The results of this detailed analysis are leveraged in §4 to identify a
single set of preferred pre-flyby periods for comparison to the post-flyby spin
35
110Pv1 (hr)110Pv2 (hr)0.050.100.150.200.250.30RMS (mag)110 110 state.
Other than limiting the Fourier coefficients to < 0.5 magnitudes in our fitting
algorithm (§3.1), the net amplitudes of the pre-flyby fits were not restricted in
any way. However, due to the large temporal gaps in the pre-flyby data, this
resulted in some cases where the model amplitudes were much greater than
the well-defined post-flyby amplitude of 1.9 magnitudes (Figure 4). Without
detailed knowledge of Duende's shape, which we don't currently have, we
can not make a clear estimate for the expected amplitude at the time of
the pre-flyby viewing geometry. Instead, based on the ensemble of measured
amplitudes for other similarly sized NEOs, including those in non-principal
axis rotation (Warner et al., 2009), we expect Duende to have a maximum
lightcurve amplitude of less than about 2.5 magnitudes. In our pre-flyby grid
search we find that 70% of fits have amplitudes ≤ 2.5 magnitudes, suggesting
that the majority of fitted amplitudes are reasonable. The exact threshold for
when a fitted amplitude becomes unphysical remains unknown.
To provide a complementary probe of periodic signatures in the pre-flyby
data, the same FFT WindowClean analysis from §3.1 was applied to these
data (Figure 10). Though the sparseness and low quality of the pre-flyby data
results in a much noisier FFT, this result does display different peaks than the
post-flyby case. With the understanding that additional physically meaningful
peaks could be obscured by noise, there are useful comparisons between these
FFT peaks and the periods identified in the RMS fitting analysis. The most
prominent peaks in Figure 10 are at frequencies around f(cid:48)
f(cid:48)
c = 7.2, and f(cid:48)
d = 11.7 cycles per day (Table 5). The second highest peak
around 11.7 cycles per day is a possible counterpart to the series of ∼ 4.1
hour solutions in the RMS fits (2 / 4.1 hr = 11.7 day−1). The largest peak in
a = 2.0, f(cid:48)
b = 2.9,
36
Fig. 10. FFT power spectra of the pre-flyby lightcurve data. The locations of the 4
most prominent peaks in the post-flyby FFT are shown as dashed grey lines. The
cleaned power spectrum (bottom) shows several peaks (f(cid:48)
a−d) which are discussed
in the text. Several of these can be mapped to periods identified in our RMS fitting
analysis. In general the peaks here do not match those in the post-flyby FFT.
this FFT occurs around 7.2 cycles per day. This is close to the peak associated
with the second order linear combination of our preferred pre-flyby period pair
of P (cid:48)
around 6 hours. The peak at 2.9 cycles per day is close to 1/P (cid:48)
2 = 24.2 (§4). It also is close to 2/P (cid:48)
1 = 8.37 and P (cid:48)
1 for period solutions
1 for periods
around 8 hours. The peak centered on 2 cycles per day may correspond to
2/P (cid:48)
2 for solutions around 24 hours.
In general the FFT for the pre-flyby data displays different peaks than that
of the post-flyby case (Figure 6). However, there are some similarities, par-
ticularly for our preferred period solutions. Both pre and post-flyby power
spectra show peaks associated with periods around 24 hours (2/24.2 hr for
the former, 1/23.7 hr for the latter), and with periods around 8 hours (1/8.37
37
051015cycles / day0.0000.0050.0100.015Powerdirty051015cycles / day0.0000.0010.0020.0030.004Powercleanfvafvbfvcfvd1, P (cid:48)
P (cid:48)
(hr)
2
4.5, 16.4
6.26, 24.0
6.37, 22.8
6.84, 24.0
7.56, 23.9
RMS
(mag)
0.06
0.09
0.09
0.10
0.10
8.37, 24.2 0.10
f(cid:48)
a
f(cid:48)
b
(2.0 day−1)
(2.9 day−1)
-
2/P(cid:48)
2/P(cid:48)
2/P(cid:48)
2/P(cid:48)
2/P(cid:48)
2
2
2
2
2
1/P(cid:48)
1/P(cid:48)
2
2
2
2/P(cid:48)
1 - 1/P(cid:48)
1 - 1/P(cid:48)
-
1/P(cid:48)
1/P(cid:48)
1
1
f(cid:48)
c
2
(7.2 day−1)
1/P(cid:48)
1 + 1/P(cid:48)
2/P(cid:48)
2/P(cid:48)
2/P(cid:48)
1
1
1
-
2/P(cid:48)
1 + 2/P(cid:48)
2
f(cid:48)
d
(11.7 day−1)
2/P(cid:48)
3/P(cid:48)
1
1
-
-
-
4/P(cid:48)
1
Table 5
Selection of pre-flyby period solutions. Our preferred solution is highlighted.
hr for the former, 1/8.71 hr for the latter). The key question that remains is
whether these differences, including the different best-fit periods in our RMS
analysis, are due to noise in the pre-flyby data, are due to changes in observing
geometry pre- and post-flyby, and/or indicate a change in rotation state due
to tidal effects during the encounter. We address this question in the follow-
ing section, noting that the more comprehensive analysis of the solid body
dynamics presented in Benson et al. (2019) provides results that complement
arguments presented here. We note here an important point that generalized
models for bodies in NPA rotation find that viewing geometry changes can
modify the strength of FFT peaks, but typically do not cause a shift of the
peaks in frequency space (Samarasinha & Mueller, 2015).
4 Pre/Post-flyby Comparison
Our initial objective for the lightcurve investigation was to probe for any de-
tectable changes in the rotation state of the asteroid. While the pre-flyby data
are not of sufficient quality to definitively address this goal, we do compare
analyses of the pre and post-flyby data to identify preferred pre-flyby periods
38
and to address the plausibility that Duende experienced rotational changes
during the planetary encounter. We investigate in detail the low RMS solu-
tions returned in our least squares analysis (§4.1) and the implications of our
FFT spectral analysis (§4.2).
A brief overview of short-axis non-principal axis rotation facilitates the follow-
ing discussion (Figure 11). More detailed descriptions of the SAM coordinate
system, rotation mode, and justification for its relevance to Duende, are pro-
vided in Benson et al. (2019). A SAM state can be described by rotation about
several axes. The long axis of the body or the symmetry axis (ψ) rotates with
an average angular precessionary period P ¯φ about the net angular momentum
vector (cid:126)H. The symmetry axis also nods with a period Pθ around the nutation
angle θ. For for the long axis convention shown here, Pθ = Pψ (Benson et
al., 2019). Lastly, the body oscillates around the symmetry axis with angular
amplitude less than 90◦ and a period equal to Pψ. This description is based
on Greenwood (1987) and follows the notation of Pravec et al. (2005).
4.1 RMS Fitting
Attempts to fit the pre-flyby data with periods equal to those measured post-
flyby resulted in RMS residuals that were a factor of at least two higher than
the best fit RMS values of ∼ 0.06 magnitudes (Table 5). Forcing fits to the
pre-flyby data with the post-flyby periods in Table 4 resulted in RMS values of
0.12 for P (cid:48)
hr, and 0.12 for P (cid:48)
2 = 8.73 hr, of 0.19 for P (cid:48)
2 = 23.7 hr. With mean residuals this
1 = 8.71 hr and P (cid:48)
2 = 23.7
1 = 6.36 hr and P (cid:48)
1 = 6.37 hr and P (cid:48)
high (> 0.1 magnitude) there are clear discrepancies between the data and
the fit, in some cases with the fits not tracking clear trends (e.g. increasing or
decreasing magnitude) in the data.
39
Fig. 11. Rotational dynamics of an ellipsoid in non-principal, short-axis mode (SAM)
rotation. Duende is roughly a prolate ellipsoid 20 meters along its short axes and
40 meters along its major axis. SAM rotation can be described by oscillation of a
body around its axis of symmetry (ψ) and rotation of the symmetry axis about the
angular momentum vector ( (cid:126)H). The rotation periods associated with these two axes
are Pψ and P ¯φ respectively. Based on constraints from the radar observations and
the solid body analysis in Benson et al. (2019), the nutation angle θ is expected to
be large, ∼ 70 − 85◦. For axis convention shown here, the nodding rate Pθ = Pψ.
Based on an analysis of the solid body dynamics (Benson et al., 2019), the
post-flyby period pair with the second lowest RMS, P1 = 8.71 hr and P2 = 23.7
(Table 4), offers the most plausible description of the post-flyby rotation state.
Furthermore, this dynamical analysis suggests that P1 = Pφ and P2 = Pψ. We
expect pre-flyby solutions to display similar values of Pψ due to the difficulty
in tidally inducing period changes about the symmetry axis. Considering the
uncertainties associated with the pre-flyby data, we suggest that pre-flyby
periods within ∼ 10% of Pψ =23.7 hours would be broadly consistent with a
constant Pψ throughout the flyby. With this as a rough guideline we identify
a number of P (cid:48)
9 -- for which a low RMS is realized. This box encloses a range of P (cid:48)
2 period solutions -- enclosed by the white box in Figure
1 values
1 and P (cid:48)
40
from about 2 hours up to 8.4 hours, and P (cid:48)
Although these low RMS solutions go down to P (cid:48)
2 values from about 21 to 26 hours.
1 as small as 2 hours, a suite of
numerical models looking at the outcome of randomized encounter geometries
find that the smallest possible pre-flyby P ¯φ is about 6 hours assuming a post-
flyby value of 8.71 hr (Figure 10 in Benson et al., 2019). We thus set a boundary
at P (cid:48)
1 = 6 hours, anticipating more modest period changes in line with the most
likely simulated outcomes (Benson et al., 2019), and then search for best fit
solutions within the ranges 6 < P (cid:48)
1 < 8.4 hours and 21 < P (cid:48)
2 < 26 hours
without restricting the periods as was done in the grid search. This leads to a
set of 5 pre-flyby solutions that are close to the post-flyby periods and have
P (cid:48)
2 within 10% of 23.7 hours (Table 5). The RMS residuals for these solutions
are all equal to about 0.1 magnitude, which is not the lowest found in the
global grid search (Figure 9), but we favor these solutions as most plausible
based on the solid body dynamics (Benson et al., 2019) and based on the FFT
analysis below. Due to the sparseness and lower quality of the pre-flyby data it
is unclear whether the difference in RMS between these five solutions (∼ 0.1)
and the global minimum (=0.06) is statistically significant. Nevertheless, it is
interesting to consider these solutions in slightly more detail. All five of these
pre-flyby solutions centered around P (cid:48)
P (cid:48)
1 than the preferred post flyby P1 = 8.7 hours. If we thus interpret these in
2 = 23.7 hr are at significantly smaller
the same manner as the post-flyby periods, then this suggests the intriguing
possibility that Duende's rotation slowed down, i.e. Pφ increased during the
planetary flyby. The specific pre-flyby solution of P (cid:48)
1 = 8.37 hr and P (cid:48)
2 = 24.18
hr (Table 5) falls nearest the most likely outcome of these numerical models.
At first glance, it would seem that the pre-flyby solution of 6.37 and 22.78
hours (Table 5) could have gone unchanged to the post-flyby solution of 6.37
and 23.7 hours (Table 4). However, this scenario seems inconsistent with the
41
delay-Doppler radar constraint requiring P ¯φ ∼ 8 hours. As such, we suggest
that the RMS fits to the data suggest a small increase in P ¯φ from 8.37 hr
pre-flyby to 8.71 hr post-flyby.
4.2 FFT Power Spectra
We also employ the FFT analysis as a means to further probe the pre and
post-flyby rotation states. The FFT power spectra alone (Figures 6 and 10)
unfortunately do not provide conclusive evidence for a change in rotation
state. The single strongest peaks in each of these FFTs are roughly in the same
location: 7.6 cycles per day for the post-flyby case and 7.2 cycles per day for the
pre-flyby case. For our preferred post-flyby solution of P ¯φ = 8.71 hr and Pψ =
23.7 hr, these would correspond to the second order linear combination 2/P ¯φ
+ 2/Pψ. For these periods the pre-flyby FFT peaks around 2 and 2.9 could
correspond to 2/Pψ and 1/P ¯φ respectively. However, the complete absence of
the prominent post-flyby peaks at 3.78 and 5.46 cycles per day in Figure 10
suggests that the pre-flyby data either sample a different rotation state or are
too sparse to efficiently probe all of the frequencies present in the post-flyby
FFT. It is expected that if Duende's rotation state did not change, then the
frequencies of the FFT peaks should remain fixed while the relative intensities
of the peaks might change (Samarasinha & Mueller, 2015).
To address whether these differences between the two FFTs are significant, we
run a simple Monte Carlo experiment to re-sample the post-flyby lightcurve
with the sparse cadence of the pre-flyby observations. This test will address
two specific objectives: (1) determine whether such sparse sampling can accu-
rately reproduce the major peaks retrieved from denser observations, and (2)
determine the likelihood of whether the different peaks in the noisy pre-flyby
42
FFT could simply be due to sparse sampling of an unchanged post-flyby ro-
tation state. We run this test on the model fit to the post-flyby data so as to
have a noise-free representation of the lightcurve of a non-principal axis rota-
tor. This model is extrapolated out for 7 days relative to time zero defined by
the start of the post-flyby data set. We repeatedly sample this ideal lightcurve
at specific times that exactly mimic the pre-flyby observations. For each sam-
ple set we randomly assign from a uniform distribution a 0 to 0.5 day offset
relative to time zero, and we assign from a gaussian distribution with a full-
width-half-maximum = 0.1 magnitude random noise to each sample point. For
each sample set we run the WindowClean analysis (§3.1) and record the five
peaks with the highest power in the cleaned FFT. This sampling procedure is
repeated 1000 times.
To address the first objective, we first consider the fraction of the 1000 trials
that reproduce the "true" FFT peaks. Specifically we look for whether the
sparse sampling can retrieve the two major post-flyby FFT peaks at 7.6 and
5.5 cycles per day (Figure 12). We consider a peak to match one of these
frequencies if it comes within 0.15 cycles / day, roughly the full-width-half-
maximum of the peaks in the pre-flyby FFT power spectrum (Figure 10). We
find that at least one of the these peaks is retrieved in 48% of trials, that both
of them are present in 12% of trials, and that neither of them shows up in 40%
of trials. Therefore, in a slight majority (60%) of cases the sparse sampling does
retrieve accurate information about at least one of the frequencies present in
the underlying lightcurve. This suggests that the pre-flyby FFT is more likely
that not to contain physically meaningful information about rotation state.
Since it is suggestive that the sparse sampling can retrieve physically meaning-
ful frequencies, we now address the likelihood that the peaks in the pre-flyby
43
Fig. 12. Fraction of 1000 Monte Carlo trials in which the specified FFT peaks
are retrieved when sparsely sampling a model of the post-flyby lightcurve at a
cadence that mimics the pre-flyby observations. The top panel represents the ability
of the sparse sampling scheme to retrieve the "correct" solution, in this case the 7.6
cycle/day and/or the 5.5 cycle/day primary peaks from the post-flyby FFT. The
sparse sampling retrieves one of these peaks in 48% of trials, both of them in 12%
of trials, and neither in 40% of trials. The bottom panel represents the likelihood
that the primary peaks seen in the pre-flyby FFT are an artifact of sparse sampling.
This shows it is unlikely that sparse sampling alone could result in an FFT with
peaks at 7.3 and 11.8 cycles per day, thereby suggesting that a change in rotation
state may have occurred.
FFT could be a spurious outcome of noise and under-sampling. The two largest
peaks in the pre-flyby FFT are at 7.3 and 11.8 cycles per day (Figure 10). If
the rotation state went unchanged during the flyby, then we would expect
sparse sampling of the post-flyby model to retrieve at least one of those peaks
in a majority (∼ 60%) of trials. This is not the case (Figure 12). In nearly 80%
of our trials neither the 7.3 nor the 11.8 peak were detected. Again a threshold
of 0.15 cycles / day was used to determine whether a trial peak matched the
7.3 and 11.8 pre-flyby frequencies. In only 1 out of the 1000 trials did the
44
7.6 or 5.57.6 and 5.5Neither0.00.20.40.60.8Fraction of trialsPost-flyby peaks7.3 or 11.87.3 and 11.8NeitherFFT peaks in sparsely sampled lightcurve (cycles/day)0.00.20.40.60.81.0Fraction of trialsPre-flyby peakssparse sampling of the post-flyby model find peaks at both 7.3 and 11.8 cycles
per day.
The combination of these results suggests that if the rotation state of Duende
went unchanged, then the sparse sampling during the pre-flyby epoch was
more likely than not to retrieve within our threshold of 0.15 cycles per day
at least one of the major FFT peaks seen in the densely observed post-flyby
case. Of course the marked viewing geometry changes associated with our
observations (Figure 3) are a likely cause of different lightcurve morphologies
pre and post-flyby. However, models suggest that such changes would affect
the relative strength of FFT peaks and not cause peaks to shift significantly in
frequency space (Samarasinha & Mueller, 2015). Thus the differences between
the pre and post flyby FFT power spectra are unlikely to be attributable to
sparse observational sampling or viewing geometry changes. If in fact these
differences are physically meaningful, then this offers an alternative analysis
to the RMS fitting that also suggests Duende may have experienced rotational
changes during the flyby.
5 Summary and Discussion
We have presented data from an extensive observing campaign of visible
wavelength spectroscopy and photometry focused on the near-Earth aster-
oid 367943 Duende. Observations were conducted at the time of this object's
discovery in 2012, and surrounding its near-Earth flyby at a distance of 4.2
Earth radii on 15 February 2013. In general the observations were focused
on monitoring the object through the flyby to search for evidence of physical
changes induced by gravitational interactions with the Earth. The spectro-
45
scopic observations suggest an L-, K-, or Ld- spectral type, but did not reveal
any spectral changes within the systematic uncertainty of the data. Lightcurve
photometry revealed that Duende was clearly in a non-principal axis rotation
state both prior to and following the planetary encounter. The multi-hour
rotation rate makes Duende the slowest known non-principal axis rotator for
asteroids smaller than 100 meters. Such an excited rotation state could have
been induced during prior planetary encounters (Scheeres et al., 2000). For an
object with such a low Earth encounter distance, it is unlikely that the 2013
flyby was the first time it experienced close passage to the Earth.
It is not always possible to assign the rotation rates retrieved from period
analysis techniques to physical rotation axes of a body. However, the thorough
coverage of our post-flyby lightcurve coupled with delay-doppler radar imaging
and inertial constraints (Benson et al., 2019) strongly suggests that Duende is
now in a non-principal, short-axis mode rotation state with a body-rotation
period Pψ = 23.67 hr and a precession period P ¯φ = 8.71 hr. Dynamical con-
siderations by Benson et al. (2019) suggest that any changes in rotation state
would have been largest around the precession axis φ. Simulations performed
by these authors suggest that Duende could have experienced a wide range
of possible changes in rotation state during the flyby (e.g. P ¯φ changing any-
where from -25% up to +200%), with the most likely outcome a more modest
percent-level increase in P ¯φ.
Our pre-flyby lightcurve was much less extensive and resulted in looser con-
straints on the pre-encounter rotation state. Leveraging what we learned from
the post-flyby data, we provide a series of plausibility arguments that suggests
a spin down scenario is consistent with the data. Our preferred solution based
on fits to the data, Fourier analyses, and dynamical arguments suggests that
46
P ¯φ changed from 8.4 to 8.7 hours and that Pψ remained roughly unchanged
at a rate around 24 hours. This result is consistent with that of Benson et al.
(2019) where more detailed dynamics and generation of synthetic lightcurves
were considered. However, the sparse temporal coverage and lower signal-to-
noise of the pre-flyby photometry ultimately made it difficult to definitively
determine whether a spin down did in fact happen.
The results from this campaign highlight the need for extensive data coverage
on both sides of planetary encounter. The slow multi-hour and non-principal
axis rotation state of Duende made it particularly challenging to collect suf-
ficient lightcurve photometry, though such a rotation state, in comparison to
a fast principal axis rotator, would enhance any gravitationally induced ef-
fects during planetary encounters (Scheeres et al., 2000). Our analyses of the
lightcurves suggest that the analytic tools we employed (functional fitting,
FFT) are sufficient to extract detailed information about solid-body rotation
states given data of high enough quality.
The flyby of Duende was unusual in that a full 1 year of advance notice was
available. However such planetary encounters are not necessarily rare amongst
NEOs. Since Duende flew by in 2013, eighteen other NEOs have passed by
the Earth at less than 0.1 lunar distance or 6 Earth radii (though encounters
with objects of Duende's size or bigger are less common). Such future plane-
tary flybys, given enough advance notice, will provide opportunities to further
test the phenomena explored here. Most notably, the 2029 encounter of the
approximately 300-meter asteroid 99942 Apophis at a distance <6 Earth radii
will be an excellent opportunity to test models predicting significant changes
in that asteroid's rotation state (Scheeres et al., 2005). Observations of such
flybys will further inform models for the currently unknown internal structure
47
and cohesive properties of asteroids in the near-Earth population. As ongoing
sky surveys continue to increase the annual yield of newly discovered near-
Earth objects (e.g. Christensen et al., 2012; Tonry, 2011), the rate of advance
predictions of planetary flybys is increasing. This rate will grow even further
as next generation surveys like the Large Synoptic Survey Telescope (LSST;
Jones et al., 2009) come online. Experiments, like what we have presented
here, to investigate processes of planetary geophysics in real-time will thus
become increasingly feasible.
Acknowledgements
We are extremely grateful to Marina Brozovi´c and an anonymous reviewer
for careful reading and thoughtful insights that significantly improved this
manuscript. N.M. would like to acknowledge support from the National Science
Foundation Astronomy and Astrophysics Postdoctoral fellowship program and
the Carnegie Institution for Science, Department of Terrestrial Magnetism
during the initial planning and executions stages of this campaign. Com-
pletion of the work was supported by NASA grant numbers NNX14AN82G
and NNX17AH06G (PI N. Moskovitz) issued through the Near-Earth Object
Observations program to the Mission Accessible Near-Earth Object Survey
(MANOS). D.P is grateful to the AXA research fund for their postdoctoral fel-
lowship. F.M was supported by the National Science Foundation under Grant
Number 1743015. The observations presented here would not have been pos-
sible without the outstanding contributions from numerous telescope opera-
tors and support scientists, including David Summers, Daniel Harbeck and
Ralf Kotulla at the WIYN 3.5m Observatory, Diane Harmer at the Kitt Peak
2.1m, Shai Kaspi at the Wise Observatory in Israel, and Chad Trujillo at
48
Gemini. We are grateful to Lance Benner (JPL) and Michael Busch (NRAO)
for helping guide us through the interpretation of their radar data. Beatrice
Mueller kindly provided code for the WindowClean computation. Despite our
efforts a number of planned observations were unsuccessful, we would like to
acknowledge the efforts to observe made by Michelle Bannister, Ovidiu Vadu-
vescu, and Herve Bouy. We appreciate Bruce Gary making his lightcurve data
publicly available. None of this work would have been possible without the
invaluable resources provided by the Minor Planet Center and the JPL Hori-
zons system. A. Stark is grateful for training received at the House of Black
and White in the free city of Braavos. The paper is based in part on data ob-
tained at Kitt Peak National Obseratory with the WIYN observatory and the
Kitt Peak 2.1m. Kitt Peak National Observatory, National Optical Astron-
omy Observatory is operated by the Association of Universities for Research
in Astronomy (AURA) under cooperative agreement with the National Sci-
ence Foundation. The WIYN observatory is a join facility of the University
of Wisconsin-Madison, Indiana University, Yale University, and the National
Optical Astronomy Observatory. Data were also obtained with the VATT:
the Alice P. Lennon Telescope and the Thomas J. Bannan Astrophysics Fa-
cility. This paper also includes observations obtained at Gemini-South Ob-
servatory, which is operated by the Association of Universities for Research
in Astronomy, Inc., under a cooperative agreement with the NSF on behalf
of the Gemini partnership: the National Science Foundation (United States),
National Research Council (Canada), CONICYT (Chile), Ministerio de Cien-
cia, Tecnolog´ıa e Innovaci´on Productiva (Argentina), Minist´erio da Ciencia,
Tecnologia e Inova¸cao (Brazil), and Korea Astronomy and Space Science Insti-
tute (Republic of Korea). Gemini data were processed using the Gemini IRAF
package. The data presented here were also obtained [in part] with ALFOSC,
49
which is provided by the Instituto de Astrofisica de Andalucia (IAA) under
a joint agreement with the University of Copenhagen and NOTSA. The AL-
FOSC observations were made with the Nordic Optical Telescope, operated
by the Nordic Optical Telescope Scientific Association at the Observatorio del
Roque de los Muchachos, La Palma, Spain, of the Instituto de Astrofisica de
Canarias. This paper also uses observations made at the South African As-
tronomical Observatory (SAAO). Finally, this paper includes data gathered
with the 100" Irenee du Pont telescope and the 6.5m Magellan Baade tele-
scope, both located at Las Campanas Observatory, Chile and operated by the
Carnegie Institution for Science.
References
Bell, J. 1988, A probable asteroidal parent body for the CO or CV chondrites.
Meteoritics 23, 256-257.
Belton, M. and Gandhi, A. 1988, Application of the CLEAN algorithm to
cometary light curves. Bull. Am. Astron. Soc. 20, 836.
Benner L., Brozovic, M., Giorgini, J. D., Jao, J. S., Lee, C. G, Busch, M. W.,
and Slade, M. A. 2013, Goldstone radar images of near-Earth asteroid 2012
DA14. American Astronomical Society, DPS meeting #45, 101.02.
Benson, C. J., Scheeres, D., Moskovitz, N. 2019, Spin State Evolution of As-
teroid (367943) Duende During its 2013 Earth Flyby. Submitted to Icarus.
Bertin, E. and Arnouts, S. 1996, SExtractor: Software for source extraction.
A&AS 117, 393-404.
Binzel, R. P. and 9 co-authors. 2010, Earth encounters as the origin of fresh
surfaces on near-Earth asteroids. Nature 463, 331-334.
Bottke, W. F. and Melosh, J. H., 1996, Binary Asteroids and the Formation
50
of Doublet Craters. Icarus 124, 372-391.
Brosch, N., Polishook, D., Shporer, A., Kaspi, S., Berwald, A., and Manulis,
I. 2008, The Centurion 18 telescope of the Wise Observatory. Astrophysics
and Space Science 314, 163-176.
Brown, T.M. and 54 co-authors. 2013, Las Cumbres Observatory Global Tele-
scope Network. PASP 125, 1031-1055.
Brozovi´c, M. and 19 co-authors. 2017, Goldstone radar evidence for short-axis
mode non-principal-axis rotation of near-Earth asteroid (214869) 2007 PA8.
Icarus 286, 314-329.
Bus, S. J. and Binzel, R. P. 2002, Phase II of the Small Main-Belt Asteroid
Spectroscopic Survey: A Feature-Based Taxonomy. Icarus 158, 146-177.
Burns, J. A. and Safronov, V. S. 1973, Asteroid nutation angles. MNRAS 165,
403-411.
Christensen, E. and 8 co-authors. 2012, The Catalina Sky Survey: Current and
Future Work. American Astronomical Society, DPS meeting #44, 210.13.
Dandy, C. L., Fitzsimmons, A. and Collander-Brown, S. J. 2003, Optical colors
of 56 near-Earth objects: trends with size and orbit. Icarus 163, 363-373.
Dressler, A. and 14 co-authors. 2011, IMACS: The Inamori-Magellan Areal
Camera and Spectrograph on Magellan-Baade. PASP 123, 288-332.
de Leon, J. and 11 co-authors. 2013, Visible and near-infrared observations of
asteroid 2012 DA14 during its closest approach of February 15, 2013. A&A
555, L2.
Farnocchia, D., Chesley, S. R., Brown, P., G. and Chodas, P. W. 2016, The
trajectory and atmospheric impact of asteroid 2014 AA. Icarus 274, 327-333.
Gary, B. L. 2013, Asteroid 2012 DA14 Rotation lightcurve. Minor Planet Bul-
letin 40, 122-124.
Galache, J. L., Beeson, C. L., McLeod, K. K., and Elvis, M. 2015, The need
51
for speed in Near-Earth Asteroid characterization. Planet. Space Sci. 111,
155-166.
Greenwood, D. 1987, Principles of Dynamics, Prentice Hall: Englewood Cliffs,
N.J., 2nd Edition.
Hapke, B. 2001, Space weathering from mercury to the asteroid belt. JGR
106, 10039-10073.
Harbeck, D. R. and 16 co-authors. 2014, The WIYN one degree imager 2014:
performance of the partially populated focal plane and instrument upgrade
path. Proceedings of the SPIE, Volume 9147, id. 91470P.
Harris, A. 1994, Tumbling asteroids. Icarus 107, 209-211.
Hirabayashi, M., Scheeres, D. J., and Holsapple, K. A. 2013, Constraints on
the size of asteroid 216 Kleopatra using internal stresses. 44th Lunar and
Planetary Science Conference, 1719, 1592.
Holsapple, K. A. 2004, Equilibrium figures of spinning bodies with self-gravity.
Icarus 172, 272-303.
Hudson, R. S. and Ostro, S. J. 1995, Shape and non-principal axis spin state
of asteroid 4179 Toutatis. Science 270, 84-86.
Jenniskens, P. and 34 co-authors. 2009, The impact and recovery of asteroid
2008 TC3. Nature 458, 485-488.
Jones, R. L. and 9 co-authors. 2009, Solar System science with LSST. Earth,
Moon and Planets 105, 101-105.
Kaasalainen, M. 2001, Interpretation of lightcurves of precessing asteroids.
A&A 376, 302-309.
Kryszczynska, A., Kwiatkowski, T., Breiter, S. and Michalowski, T. 1999,
Relation between rotation and lightcurve of 4179 Toutatis. A&A 345, 643-
645.
Lantz, B. and 14 co-authors. 2004, SNIFS: A wideband integral field spectro-
52
graph with microlens array. Proc. SPIE 5249, 146-155.
Lantz, C., Clark, B. E., Barucci, M. A., and Lauretta, D. S. 2013, Evidence for
the effects of space weathering spectral signatures on low albedo asteroids.
A&A 554, A138.
Monet, D. and 28 co-authors. 2003, The USNO-B Catalog. AJ 125, 984-993.
Moskovitz, N. 2012, Colors of dynamically associated asteroid pairs. Icarus
221, 63-71.
Moskovitz, N., Prieto, J., Sheppard, S. S., Williams, G. V., and Marsden, C. L.
2013, 2012 DA14. Minor Planet Electronic Circular, 2013-C19.
Mueller, B., Samarasinha, N. H. and Belton, M. J. S. 2002, The Diagnosis of
Complex Rotation in the Lightcurve of 4179 Toutatis and Potential Appli-
cations to Other Asteroids and Bare Cometary Nuclei. Icarus 158, 305-311.
Nesvorny, D., Jedicke, R., Whiteley, R.J., and Ivezic, Z., 2005, Evidence for
asteroid space weathering from the Sloan Digital Sky Survey. Icarus 173,
132-152.
Ostro, S. J. and 13 co-authors. 1995, Radar images of asteroid 4179 Toutatis.
Science 270, 80-83.
Polishook, D. and Brosch, N. 2009, Photometry and spin rate distribution of
small-sizes main belt asteroids. Icarus 199, 319-332.
Pravec, P. and 19 co-authors. 2005, Tumbling Asteroids. Icarus 173, 108-131.
Pravec, P. and 19 co-authors. 2014, The tumbling spin state of (99942)
Apophis. Icarus 233, 48-60.
Richardson, D. C, Bottke, W. F. and Love, S. G., 1998, Tidal Distortion and
Disruption of Earth-Crossing Asteroids. Icarus 134, 47-76.
Roberts, D. H., Lehar, J., and Dreher, J. W. 1987, Time-series analysis with
CLEAN. I. Derivation of a spectrum. AJ 93, 968-989.
Roche, E., 1849, La figure d'une masses fluide soumise a l'attraction d'un point
53
eloigne (The figure of a fluid mass subjected to the attraction of a distant
point). Academie des Sciences de Montpellier: Memoirs de la section de
sciences 1, 243-262.
Roeser, S., Demleitner, M., and Schilbach, E. 2010, The PPMXL Catalog of
Positions and Proper Motions on the ICRS. Combining USNO-B1.0 and the
Two Micron All Sky Survey (2MASS). AJ 139, 2440-2447.
Samarasinha, N. H. and Mueller, B. 2015, Component periods of non-
principal-axis rotation and their manifestations in the lightcurves of as-
teroids and bare cometary nuclei. Icarus 248, 347-356.
Sanchez, J. A., Reddy, V., Nathues, A., Cloutis, E. A., Mann, P., and
Hiesinger, H. 2012, Phase reddening on near-Earth asteroids: Implications
for mineralogical analysis, space weathering and taxonomic classification.
Icarus 220, 36-50.
Sanchez, P. and Scheeres, D. J. 2014, The strength of regolith and rubble pile
asteroids. MAPS 49, 788-811.
Scheeres, D. J., Ostro, S. J., Werner, R. A., Asphaug, E., and Hudson, R. S.
2000, Effects of gravitational interactions on asteroid spin states. Icarus 147,
106-118.
Scheeres, D. J., Marzari, F., and Rossi, A. 2004, Evolution of NEO rotation
rates due to close encounters with Earth and Venus. Icarus 170, 312-323.
Scheeres, D. J., Benner, L. A. M., Ostro, S. J., Rossi, A., Marzari, F., and
Washabaugh, P. 2005, Abrupt alteration of asteroid 2004 MN4's spin state
during its 2029 Earth flyby. Icarus 178, 281-283.
Sonnett, S., Kleyna, J., Jedicke, R., and Masiero, J. 2011, Limits on the size
and orbit distribution of main belt comets. Icarus 215, 534-546.
Takahashi, Y., Busch, M. W. and Scheeres, D.J. 2013, Spin state and moment
of inertia characterization of 4179 Toutatis. AJ 146, 95-104.
54
Takahashi, J. and 9 co-authors. 2014, Near-infrared colors of asteroid 2012
DA14 at its closest approach to Earth: Observations with the Nishiharima
Infrared Camera (NIC). PASJ 66, 53 (1-7).
Tody, D. 1986, The IRAF Data Reduction and Analysis System. Proceedings
of the Instrumentation in Astronomy VI Meeting, Tucson, AZ. Society of
Photo-Optical Instrumentation Engineers. p 733.
Tonry, J. 2011, An early warning system for asteroid impacts. PASP 123,
58-73.
Warner, B. D., Harris, A. W., and Pravec, P. 2009, The asteroid lightcurve
database. Icarus 202, 134-146.
Yu, Y., Richardson, D. C., Michel, P., Schwartz, S. R., and Ballouz, R.-L. 2014,
Numerical predictions of surface effects during the 2029 close approach of
Asteroid 99942 Apophis. Icarus 242, 82-96.
Zacharias, N. and 29 co-authors. 2010, The Third US Naval Observatory CCD
Astrograph Catalog (UCAC3). AJ 139, 2184-2199.
55
|
0902.4018 | 1 | 0902 | 2009-02-23T21:41:51 | Search for very close approaching NEAs | [
"astro-ph.EP"
] | A simulation of de-biased population of NEAs is presented. The numerical integration of modeled orbits reveals geometrical conditions of close approaching NEAs to the Earth. The population with the absolute magnitude up to H=28 is simulated during one year. The probability of possible discoveries of the objects in the Earth vicinity is discussed. | astro-ph.EP | astro-ph |
Contrib. Astron. Obs. Skalnat´e Pleso 36, 171 -- 180, (2006)
Search for very close approaching NEAs
P. Veres, L. Kornos and J. T´oth
Department of Astronomy, Physics of the Earth and Meteorology, Faculty of
Mathematics, Physics and Informatics, Comenius University, 842 48
Bratislava, The Slovak Republic
Received: May 11, 2006; Accepted: September 6, 2006
Abstract. A simulation of de-biased population of NEAs is presented. The
numerical integration of modeled orbits reveals geometrical conditions of close
approaching NEAs to the Earth. The population with the absolute magni-
tude up to H = 28 is simulated during one year. The probability of possible
discoveries of the objects in the Earth's vicinity is discussed.
Key words: asteroid -- NEA -- meteoroid -- orbital distribution
1. Introduction
The population of small NEOs (Near Earth Objects) with the absolute mag-
nitude up to H = 28 is still not very well understood. These objects with the
diameter of about 10 m assuming albedo of 0.14 (Stuart, Binzel 2004) represent
transition objects among asteroids, comets and meteoroids.
As we mentioned in our previous paper (T´oth, Kornos 2002), discoveries of
new NEOs are influenced by strong observational selection effects. We are lim-
ited by sensitivity of telescopes, a rapid angular velocity at close encounters in
the sky, almost no concentration towards ecliptic - that is why new NEOs can
be found in the entire sky. Current NEO discovery programs (e.g. Catalina Sky
Survey, LINEAR, Spacewatch, NEAT, LONEOS) are primarily focused on mi-
nor planets about 1 km in diameter and larger. The strategy is to discover such
asteroids in larger geocentric distances, that means to reach the visual magni-
tude of about 20−22, instead of a larger field of view. The search programs cover
the whole sky within one lunation (approx. 20 days). The magnitude limitation
for current NEO discovery strategy exclude smaller objects in larger distances.
These small NEOs could be discovered only during close encounters with the
Earth. But the celestial objects with a rapid angular velocity are difficult to
discover by this type of telescopes. The current number of known smaller NEOs
with H > 19 decreases rapidly with increasing absolute magnitude (Fig. 1).
This paper is focused on the population of very small NEOs, up to 10 m
in diameter and their possible discoveries in the close vicinity of the Earth.
This range size of NEOs is also very dangerous for life on the Earth, as their
atmospheric blasts can produce substantive damage on the surface (e.g. Tun-
guska event in 1908). Moreover, the collisions with small objects are much more
frequent than with larger asteroids.
172
P. Veres, L. Kornos and J. T´oth
Figure 1. Histogram of the NEO population known by Jan. 5, 2006, versus the Stuart
and Binzel (2004) model (dashed curve).
We simulate the orbital distribution of the modeled Near Earth Asteroid
population to estimate the number of close encounters with the Earth within
the mean Earth-Moon (0.0026 AU) distance. In this close vicinity also a 10 m
asteroid in suitable geometrical conditions can reach visual magnitude about
+14m, which represents the limit for small telescopes with a short focal length.
We also propose a searching program which would be able to discover such small
objects in the close Earth's vicinity. The results from the program both positive
or negative and comparison with our simulated NEAs close approaches would
help to better understand a small asteroid/comets ratio on the NEO orbits.
2. The simulation of NEA population
The current known NEO population is influenced by several observational ef-
fects. Due to this fact, it is necessary to use a debiased model to estimate the
real population and frequency of encounters with the Earth.
We focused our simulation only on NEAs (Near Earth Asteroids) according
to Bottke et al. (2000), who used three NEA source regions, the 3 : 1 mean
motion resonance with Jupiter, the ν6 secular resonance and the intermediate
Mars-crossers source. They produced the NEA debiased orbital and absolute
magnitude size distributions model for H < 18, calibrated to 910 NEAs of this
size range by fitting it to a biased population of NEO discovered by Spacewatch.
We extrapolated the de-biased model of the orbital elements distribution
(Bottke et al., 2000) of NEA population up to H = 28, assuming the same de-
livery mechanism from the source regions for small NEAs. The orbital elements
a, e, i follow Bottke et al. distributions and the angular elements (argument
of perihelion, longitude of ascending node and mean anomaly) were generated
Search for very close approaching NEAs
173
randomly. To calibrate the extrapolated model, we used the NEO cumulative
size distribution model according to Stuart and Binzel (2004)
N (< H) = 10−3.88+0.39H ,
(1)
which is a most conservative approximation, compared to other authors (Bottke
et al., 2000 or Rabinowitz et al., 1994). A discrete distribution has been used
for each orbital element determination. For our purpose, a small range (bins) of
elements value have been chosen: ∆a = 5×10−6 AU, ∆e = 10−6, ∆i = 2.5×10−4
deg, ∆ω = ∆Ω = ∆M = 10−4 deg and for the absolute magnitude ∆H = 10−2.
The NEA a − e phase space condition (q < 1.3 AU and Q > 0.983 AU) was
applied for each generated object. Finally, we got 10 964 782 NEA simulated
bodies with the orbital elements and absolute magnitude distribution (Fig. 2).
Figure 2. The semimajor axis, inclination, eccentricity and absolute magnitude (size)
distributions of 10 964 782 simulated NEAs.
3. Results
We numerically integrated 10 964 782 modeled NEA orbits during one year with
the aim to study geometrical conditions during their close encounters with the
Earth. We were looking for very close encounters within the Moon distance
174
P. Veres, L. Kornos and J. T´oth
(0.0026 AU) and we were also interested in all asteroids which reached visual
brightness below +14m.
In the study for a backward integration of the orbital evolution a DE multi-
step procedure of Adams-Bashforth-Moulton's type up to 12th order, with a
variable step-width, developed by Shampine and Gordon (1975), was used. In-
put data of the positions of the Earth, the only considered perturbing body, were
obtained from the Planetary and Lunar Ephemerides DE406 prepared by the
Jet Propulsion Laboratory (Standish, 1998). This simple model is used due to a
short time interval (one year) of integration. Several tests proved that results are
consistent with all major planets incorporated. Each output file contains the Ju-
lian date, the heliocentric and geocentric distances perturbed by the Earth, the
phase angle, the apparent visual magnitude, the right ascension and declination
of the object.
As the main result of the one year simulation, 80 bodies inside the Moon's or-
bit were obtained, which implies collisional frequency with the Earth of 2.10−2
per year. Only 18 of them also reached the visual magnitude of +14m. As it
could be expected, the cumulative number of close approaches increases ap-
proximately as a quadratic function of geocentric distance (Fig. 3). However,
the count of bodies in smaller geocentric distances is higher than quadratic
dependance resulting purely from Keplerian motion. It is due to the Earth's
gravity, although the effect is minor, still enough to be recognized. Moreover,
the absolute magnitude distribution of the encounters within the Moon's orbit
prefers smaller objects (< 100 m) with the maximum at 10 m, the most numerous
objects in the NEA size distribution (Fig. 4).
The most important parameter to detect the encountered object by any
searching system is its brightness. The visual magnitude distribution of ap-
proaching objects obtained from our simulation is depicted in Fig. 5. The ma-
jority of objects within Moon's orbit reached brightnesses between +16m and
+18m. That is why the limiting magnitude of a small searching system would
be the most restrictive factor of its efficiency.
The next important detection parameter of an object is its angular velocity
in the sky. It represents the effective exposure time during which the object
remains on one pixel of a CCD detector. The angular velocity of the object
depends on its geocentric distance and geocentric velocity vector. The angular
velocities of simulated objects within the Moon's orbit are up to 100 arcmin/min
(Fig. 6), but the most frequent velocities are in the range of 10 − 20 arcmin/min.
The right plot of Fig. 7 shows all 80 close encounters trajectories within the
Moon's orbit projected on the sky in equatorial coordinates. There is no no-
ticeable concentration toward the ecliptic. Some objects changed their positions
over 100 degrees in one day. At the same time the phase angles changed very
quickly, which strongly influenced the visual magnitudes of simulated objects.
On the left plot of Fig. 7 there are depicted only objects inside the Moon's orbit
brighter than +14m.
Search for very close approaching NEAs
175
100
N
10
1
104
105
D (km)
106
Figure 3. Cumulative annual NEAs flux inside the Moon's orbit (solid), theoretical
quadratic dependance (dashed), D is the minimum geocentric distance. One simulated
object collided with the Earth.
diameter (m)
100
10
100
10
N
1
22
24
H
26
28
Figure 4. Cumulative annual NEAs flux inside the Moon's orbit. Size distribution.
Also, there was a question how many of simulated objects could be detected
by small telescopes in one year. Totally 119 bodies reached the visual magnitude
of +14m and brighter. Just 18 of them, as we have already mentioned, were
at the minimum encounter distance within the Moon's orbit. Other bodies,
mostly larger objects, reached +14m in greater geocentric distances at various
phase angles. There are 52 of them bright enough due to their small heliocentric
distances. There is a possibility to search for such objects at small elongations
from the Sun (mornings and evenings) or by the LASCO coronograph on SOHO
space mission.
176
P. Veres, L. Kornos and J. T´oth
Figure 5. Annual flux of NEAs inside the Moon's orbit and maximum visual magni-
tude distribution at the closest encounters.
Figure 6. Histogram of angular velocity of NEAs at the closest approaches to the
Earth within the Moon's orbit.
3.1. Possible detections
We have applied several geometrical conditions and limitations on close encoun-
ters obtained from the simulation in relation to a proposed searching system with
the limiting magnitude of +14m and weather conditions for Central Europe. In
the estimate we took into account only objects with the absolute magnitude
H > 19, as we have focused on a search for small, mostly unknown objects,
which would not be discovered by the current searching programs (see Fig. 1).
The visual magnitude of an object strongly depends on the phase angle. The
brightness at the phase angle P h = 60◦ is about 2 magnitudes lower than that
at the opposition point with P h = 0◦ in the same heliocentric and geocentric
Search for very close approaching NEAs
177
Figure 7. Close encounters trajectories within the Moon's orbit projected on the sky.
Simulated objects with the apparent magnitudes < 14m (left) compared to all 80 NEAs
inside the Moon's orbit (right). The solid curve is the ecliptic.
distances. During a very close encounter the phase angle of the object is almost
equal to its angle from the opposition point. That is why we suggest an effec-
tive search scan of the sky just within the area of 60◦ around the opposition.
This part of the sky represents the area of approximately 10 000◦2. Due to our
upcoming searching system limitation in the sky coverage area we suppose to
cover only ∼ 2 700◦2, which corresponds to the phase angle P h = 30◦.
The effective time which the object spends in the searching area (30◦ from
the opposition) reduces the probability of the detection. Another decrease occurs
after reducing the magnitude of each object according to its angular velocity,
which corresponds to the effective exposure time per pixel for a particular optical
CCD system. There is also taken into account the time of the encounter that
have to be in the nighttime for Central Europe.
According to these conditions we obtained 18 possible detections during one
year simulation for the Modra observatory (Comenius University, Bratislava,
Slovakia). Only 6 of them occurred inside the Moon's orbit. One simulated
body hit the Earth. The number of bodies in suitable geometrical conditions
slightly varies with a location on the Earth. The last reduction is needed due
the fact that only about 20 − 30 percent of the nights in Central Europe are in
good weather conditions.
Finally, we obtained only 3 − 5 objects, which would be detected per year
with the proposed searching system for the particular weather conditions in
Central Europe. This result is not very optimistic, but we have chosen for our
simulation the conservative estimate of the NEA population (H < 28) by Stuart
and Binzel (2004). The total NEO population estimates up to H < 28 by other
authors like Bottke et al. (2000) or Rabinowitz et al. (1994) are several times
(6 − 24 times) larger (Tab. 1).
Especially, analysis of objects inside the Moon orbit showed that a higher
limiting magnitude of a wide field survey system leads to more discoveries. As
it is clear from the visual magnitude distribution in Fig. 5, the +14m limiting
178
P. Veres, L. Kornos and J. T´oth
Table 1. The simulation results based on the NEAs (H < 28) population estimate
by Stuart and Binzel (2004). The number of NEAs per year: (a) brighter than +14m
with no distance limitation; (b) inside the Moon's orbit; (c) inside the Moon's orbit
and brighter than +14m; (d) brighter than +14m, observable in Central Europe and
H > 19; (e) possible discoveries by the proposed searching system. Presumptive results
are based on the population estimates by Bottke et al. (2000) and Rabinowitz et al.
(1994).
N annually
a) apparent mag.< 14
b) inside Moon orbit
c) inside Moon orbit and
vis. mag.< 14
d) vis. mag.< 14 and
observable in CE, H > 19
e) poss. discoveries by s.s.
Simulation Bottke et al., 2000 Rabinowitz et al., 1994
119
80
18
11
3
725
528
118
66
18
2636
1920
432
264
72
magnitude is not sufficient for the effectivity of the searching system. To detect
a majority of close encounters the +18m magnitude would be needed.
3.2. Searching system
The simulation of geometrical conditions during close approaches to the Earth
were done with the aim of possible discoveries of small asteroids in the Earth's
vicinity.
An example of such a searching system is described below (Fig. 8). The
parameters of the system with the CCD ST8 camera are as follows: the focal
length f = 150 mm, the aperture D = 180 mm, the resolution of 37.15 arcsec per
pixel, the field of view 5.27◦
×3.51◦, the exposure time of 30 s, which corresponds
to the limiting magnitude of +13.2m.
The pro/-posed wide-/-field short focal length search telescope is not suit-
able for precise astrometric observations, but is still useful to follow up detection
of close approaches. Such observation can provide a long arc with the sufficient
curvature for preliminary orbit determination (Milani, 2006), afterwards con-
firmed by our f/5.5 60 cm reflector. This kind of detection could be classified as
the original discovery.
The searching system is not finished yet and is in an early testing phase. But
the most restrictive factor of such a short focal length system is a low limiting
magnitude. The results of our simulation showed that the effective searching
system has to have the limiting magnitude up to +18m and would cover not
less than 10 000◦2 per night scan rate (Fig. 5).
Search for very close approaching NEAs
179
Figure 8. An example of a possible searching system for the close approaching NEOs.
4. Conclusions
We have extrapolated the de-biased model of NEA orbital distribution (Bottke
et al., 2000) up to H = 28 and calibrated it using the NEO cumulative size
distribution model of Stuart and Binzel (2004). We have numerically studied of
about 11 millions generated NEAs during one year. As the main result from the
simulation we obtained 80 close encounters with the Earth within the Moon's
orbit. Only 18 of these objects reached the visual magnitude of +14m.
The proposed wide-field short focal length survey system, assuming a ground-
based restrictions, observational selection effects and atmospherical conditions
could find several (∼ 3) NEAs per year. If the real population is larger (Rabi-
nowitz et al., 1994; Bottke et al., 2000), there is a possibility to find 18 − 72
objects per year. An increase of the limiting magnitude of the survey system to
+18m with preserving a wide field of view could lead to more discoveries.
An alternative way to compare our simulation results with observations is
to calculate an Earth's annual collisional frequency. The collisional frequency of
10 m objects calculated from DoD satellite (U.S. Department of Defence) obser-
vations of large bolides is 10−1 (Brown et al., 2002). Our simulation resulted in
the collisional frequency of 2.10−2 per year, which implies that the real popula-
tion should be at least 5 times more numerous than that used in our expanded
model up to H = 28 based on Stuart and Binzel (2004).
Acknowledgements. The authors thank to P. Brown and A. Gal´ad for valu-
able comments. This work was supported by the Slovak Grant Agency, Grant No.
1/3067/06.
180
References
P. Veres, L. Kornos and J. T´oth
Bottke, W.F., Jedicke, R., Morbidelli, A., Petit, J. M., Gladman, B.: 2000, Science 288,
2190
Brown, P., Spalding, R.E., ReVelle, D.O., Tagliaferri, E., Worden, S.P.: 2002, Na-
ture 420, 294
Milani, A.: 2006, personal communication.
Rabinowitz, D.L., Bowell, E., Shoemaker, E.M., Muionen, K.: 1994, in Hazard Due to
Comets and Asteroids, ed.: T. Gehrels, The University of Arizona Press, Tucson,
285
Shampine, L.F., Gordon, M.K.: 1975, Computer Solution of Ordinary Differential
Equations, Freeman and Comp., San Francisco
Standish, E.M: 1998, JPL IOM 312.F - 98 - 048
Stuart, J.S., Binzel, R.P.: 2004, Icarus 170, 295
T´oth, J., Kornos, L.: 2002, Acta Astron. Et Geophys. Univ. Comenianae XXIV, 61
|
1606.03030 | 1 | 1606 | 2016-06-09T17:20:41 | Analytical investigation of the decrease in the size of the habitable zone due to limited CO$_2$ outgassing rate | [
"astro-ph.EP"
] | The habitable zone concept is important because it focuses the scientific search for extraterrestrial life and aids the planning of future telescopes. Recent work has shown that planets near the outer edge of the habitable zone might not actually be able to stay warm and habitable if CO$_2$ outgassing rates are not large enough to maintain high CO$_2$ partial pressures against removal by silicate weathering. In this paper I use simple equations for the climate and CO$_2$ budget of a planet in the habitable zone that can capture the qualitative behavior of the system. With these equations I derive an analytical formula for an effective outer edge of the habitable zone, including limitations imposed by the CO$_2$ outgassing rate. I then show that climate cycles between a Snowball state and a warm climate are only possible beyond this limit if the weathering rate in the Snowball climate is smaller than the CO$_2$ outgassing rate (otherwise stable Snowball states result). I derive an analytical solution for the climate cycles including a formula for their period in this limit. This work allows us to explore the qualitative effects of weathering processes on the effective outer edge of the habitable zone, which is important because weathering parameterizations are uncertain. | astro-ph.EP | astro-ph | Analytical investigation of the decrease in the size of the
habitable zone due to limited CO2 outgassing rate
Dorian S. Abbot1
[email protected]
ABSTRACT
The habitable zone concept is important because it focuses the scientific
search for extraterrestrial life and aids the planning of future telescopes. Recent
work has shown that planets near the outer edge of the habitable zone might
not actually be able to stay warm and habitable if CO2 outgassing rates are not
large enough to maintain high CO2 partial pressures against removal by silicate
weathering. In this paper I use simple equations for the climate and CO2 budget
of a planet in the habitable zone that can capture the qualitative behavior of the
system. With these equations I derive an analytical formula for an effective outer
edge of the habitable zone, including limitations imposed by the CO2 outgassing
rate. I then show that climate cycles between a Snowball state and a warm cli-
mate are only possible beyond this limit if the weathering rate in the Snowball
climate is smaller than the CO2 outgassing rate (otherwise stable Snowball states
result). I derive an analytical solution for the climate cycles including a formula
for their period in this limit. This work allows us to explore the qualitative effects
of weathering processes on the effective outer edge of the habitable zone, which
is important because weathering parameterizations are uncertain.
Subject headings: planets and satellites: atmospheres, astrobiology
6
1
0
2
n
u
J
9
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
3
0
3
0
.
6
0
6
1
:
v
i
X
r
a
1.
Introduction
The habitable zone is defined as the region around a star where a planet with CO2
and H2O as its main greenhouse gases can support liquid water at its surface (Kasting et al.
1993). The habitable zone is relatively wide because of the silicate-weathering feedback
(Walker et al. 1981). Silicate-weathering is a geological process that removes CO2 from the
1Department of the Geophysical Sciences, University of Chicago, 5734 South Ellis Avenue, Chicago, IL
60637
-- 2 --
atmosphere and a negative (stabilizing) feedback is possible because this process is believed
to run faster at higher temperatures and higher CO2 partial pressures (Pierrehumbert 2010).
The inner edge of the habitable zone is set by the moist or runaway greenhouse (Kasting
1988; Nakajima et al. 1992; Goldblatt & Watson 2012), which should not be influenced by
the details of the silicate-weathering feedback since the CO2 should have been drawn down
to low levels when they occur. On the other hand, calculations of the outer edge of the
habitable zone generally assume that the silicate-weathering feedback can maintain CO2
at arbitrarily high levels. The outer edge is then marked by some threshold where adding
CO2 to the atmosphere no longer provides additional warming, for example if additional CO2
increases Rayleigh scattering more than it increases greenhouse warming (Kasting et al. 1993;
Kopparapu et al. 2013). This picture essentially considers the asymptotic limit of unlimited
CO2 outgassing capacity.
More recently it has been recognized that for finite CO2 outgassing rates the maximum
CO2 that a planet can achieve may be lower than the CO2 needed to keep that planet
habitable at the outer edge of the habitable zone (Tajika 2007; Kadoya & Tajika 2014). This
leads to what I will call the "effective outer edge of the habitable zone," which will depend
on the CO2 outgassing rate and the functioning of silicate weathering on the planet. Beyond
the effective outer edge of the habitable zone, a planet may experience cycles (Menou 2015;
Haqq-Misra et al. 2016) between a globally frozen, Snowball Earth climate (Kirschvink 1992;
Hoffman et al. 1998) and a habitable climate. Understanding and constraining the effective
outer edge of the habitable zone is critical because it determines our estimate of the fraction
of stars that host an Earth-like planet (Petigura et al. 2013; Kopparapu 2013), which is
essential for planning future telescopes that would observe such planets.
The purpose of this paper is to investigate what determines the position of the effective
outer edge of the habitable zone and what happens beyond the effective outer edge. The
behavior of a planet in this regime is largely determined by its uncertain silicate-weathering
behavior.
Instead of viewing this an an obstacle, I will exploit it to make progress on
the problem. The importance and uncertainty of weathering justifies the use of a simple
climate model since the errors associated with making the grave assumptions that such a
model entails pale in comparison to the uncertainty in weathering.
It also allows me to
use a relatively simple weathering parameterization that captures the expected qualitative
behavior. I will use the simplifying assumptions to derive useful formulae. This will allow us
to understand issues such as which processes determine whether the effective outer edge of
the habitable zone is inside the traditional outer edge of the habitable zone and how much
the parameters associated with these processes would need to be changed from our best
estimates of their values to change the qualitative behavior of the system. Moreover, I am
able to determine the conditions for climate cycles to occur beyond the effective outer edge
-- 3 --
of the habitable zone, an analytical solution for these cycles, and a formula for the period
of the cycles. Using more complicated climate and weathering models will alter quantitative
results, but is unlikely to alter the qualitative dependencies on parameters that my formulae
give. This work therefore complements recent work by Menou (2015) and Haqq-Misra et al.
(2016), who used more complicated radiative and climate models, but only explored a few
values of weathering parameters.
I will neglect sophisticated radiative (Kopparapu 2013; Goldblatt et al. 2013) and 3D
calculations (Leconte et al. 2013b,a; Yang et al. 2013, 2014; Wolf & Toon 2014, 2015) and
consider the following linearized, zero-dimensional model of planetary climate, similar to
that used by (Abbot et al. 2012):
C
dT
dt
=
S
4
(1 − α(T )) −
S0
4
(1 − αo) − a(T − T0) + b log(cid:18) P
P0(cid:19) ,
(1)
where a and b are constants, T is the global mean temperature, T0 is the temperature of
the reference state, P is the atmospheric partial pressure of CO2, P0 is the atmospheric
partial pressure of CO2 in the reference state, S is the stellar flux, S0 is the stellar flux in
the reference state, C is the heat capacity in units of J m−2 ◦C−1, α(T ) is the temperature-
dependent planetary albedo (reflectivity), and α0 is the albedo of the reference state. Table 1
contains a list model parameters and their standard values. The values I use here are drawn
from Abbot et al. (2012), Menou (2015), and Haqq-Misra et al. (2016), but it is important
to emphasize that the arguments in this paper do not depend on the exact values chosen.
I will let the albedo be specified by
α(T ) =(αw, T ≥ Ti,
αc, T < Ti,
(2)
where αw is the albedo of the warm climate state; αc is the albedo of the cold and icy, "Snow-
ball" climate state; and Ti is the temperature at which the planet transitions between the
two climate states. Equation (2) assumes a step-wise transition in albedo, which neglects the
effects of spatial resolution (Yang et al. 2012a,b,c; Voigt et al. 2011; Voigt & Abbot 2012).
Nevertheless, it allows us to make easier analytical progress and does not alter the qualitative
behavior of the system unless multiple Snowball-like climate states possible (Abbot et al.
2011; Rose 2015), which is a possibility we will not consider here. As we will see below,
introducing a smoothed albedo transition only introduces a repulsing fixed point in some
situations that the climate could not exist stably in. Finally, if we assume that the reference
climate state is warm, then α0 = αw.
-- 4 --
The partial pressure of CO2 is determined by outgassing and weathering
dP
dt
P0(cid:19)β
= V − W0ek(T −T0)(cid:18) P
,
(3)
where V is the CO2 outgassing rate, W0 is the rate of removal of CO2 from the atmosphere
by silicate weathering in the reference climate state, k is a rate constant for the increase in
weathering with temperature, and β is an exponent that determines how strongly weathering
depends on atmospheric CO2 partial pressure. West et al. (2005) used an analysis of river
catchments to estimate that k = 0.11 ± 0.04 ◦C−1 (1-σ error). Based on that study, the
plausible range for k is roughly 0 -- 0.2.
I will use a standard value of k=0.1 and vary k
over this range. β could be zero if land plants concentrate CO2 in the soil at the same
level regardless of the atmospheric CO2 concentration (Pierrehumbert 2010), and theoretical
arguments suggest β has a maximum of 1 (Berner 1994).
I will use a standard value of
β=0.5 (Berner 1994; Pierrehumbert 2010; Abbot et al. 2012; Menou 2015; Haqq-Misra et al.
2016), but consider variations within the plausible range. Depressurization caused by glacial
unloading can cause temporary increases in the CO2 outgassing rate (Huybers & Langmuir
2009). I neglect this affect and take the CO2 outgassing rate to be independent climate state
here, which is appropriate for longterm average behavior. Note also that in section 4 we will
consider W0=0 when T < Ti.
For the purposes of this paper I am assuming that weathering follows a similar param-
eterization on land and at the seafloor, but it should be understood that the weathering
behavior of an ocean planet would likely be quite different from that of a planet with an
Earth-like land fraction (Abbot et al. 2012). The weathering parameterization in Equa-
tion (3) is similar to that used by other authors (Berner 1994, 2004; Pierrehumbert 2010;
Abbot et al. 2012; Menou 2015; Haqq-Misra et al. 2016), but I have dropped the relatively
weak dependence on temperature that is often included to represent changes in precipitation
with temperature. The qualitative behavior of weathering as a function of temperature is
captured by Equation (3) without this additional complication. Given the ad hoc nature of
all weathering parameterizations, it is reasonable to choose the simplest parameterization
that gives the expected qualitative behavior of weathering processes given the objectives of
this paper.
Equations (1)-(3) define the system. The plan for analyzing them is as follows. First we
will consider the warm (habitable) state (section 2). I will set the albedo (Equation (2)) to
its warm state value, set the time derivatives in Equations (1) and (3) to zero, and solve the
system for conditions when the temperature is high enough for the warm state to exist. This
will allow us to put bounds on the existence of the warm state, that is, define an effective
outer edge of the habitable zone that may be more restrictive than the traditional outer
-- 5 --
edge. Next I will find nullclines of the system defined by Equations (1) and (3) (that is, lines
where the time derivatives equal zero), find their intersections (fixed points, or solutions of
the system), and determine the stability of these fixed points. Physically, this will reveal
that if Equation (3) is followed as is, enough weathering occurs that the system settles into a
stable Snowball state when the warm climate state ceases to exist, rather than into climate
cycles between Snowball and warm conditions. I will then show that climate cycles do occur
if we set weathering to zero in the Snowball state in Equation (3) (section 4). I will find
analytical solutions for the components of these cycles and derive an analytical formula for
their period that can predict well results from the more intricate model of (Menou 2015). I
will then discuss these results in section 5 and conclude in section 6.
Table 1: A list of the model parameters, their descriptions, and the standard values I use for
them. The parameter values in this table are taken from Abbot et al. (2012), Menou (2015),
and Haqq-Misra et al. (2016). Note that, following Haqq-Misra et al. (2016), I use a value
of W0 ten times higher than Menou (2015). I perform sensitivity analyses where I vary k
and β.
Parameter Description
S0
S
T0
P0
a
b
α0
αw
αc
Ti
C
V
W0
k
β
reference state stellar flux
stellar flux
reference state temperature
reference state partial pressure CO2
slope of planetary infrared emission with temperature
slope of planetary infrared emission with logarithm of CO2
reference state albedo
warm state albedo
cold state albedo
albedo transition temperature
planetary heat capacity
CO2 outgassing rate
reference state CO2 weathering rate
weathering-temperature rate constant
weathering-CO2 power law exponent
Standard Value
1365.0 W m−2
variable
15.0◦C
3×10−4 bars
2.0 W m−2 ◦C−1
10.0 W m−2
0.3
0.3
0.6
-10◦C
2×108 J m−2 ◦C−1
variable
70 bars Gyr−1
0.1 ◦C−1
0.5
2. Conditions on the existence of a habitable climate state
In this section we will consider the conditions that allow the carbon cycle to maintain
a planet in the warm climate state, which is necessary for the planet to be considered
-- 6 --
habitable in the traditional sense. This requires a steady-state solution, so we can set the
time derivatives in Equations (1) and (3) to zero. We can solve this system to find
Tw − T0 =
b log(cid:16) V
W0(cid:17) + β
kb + aβ
4 (S − S0)(1 − αw)
,
(4)
where Tw is the warm state temperature. We could rewrite Equation (4) in non-dimensional
form, but I will leave it, and subsequent equations, in dimensionful form to make them more
accessible physically. From Equation (4) we can see that increasing the CO2 outgassing rate
logarithmically increases the warm state temperature, and that increased outgassing warms
the warm state more if the climate is more sensitive to CO2 (higher b), provided β 6= 0. We
can also see that the influence of stellar flux on the warm state temperature depends on β,
the weathering-CO2 power-law exponent. If β = 0, then changing the stellar flux has no
effect on the warm state temperature as long as the stellar flux is high enough that Tw ≥ Ti,
so that the warm state can exist. Instead, the warm state temperature is determined by the
outgassing rate. As β increases, the warm climate state becomes increasingly sensitive to
changes in stellar flux.
We can solve for the warm state CO2 partial pressure (Pw) as follows
log(cid:18) Pw
P0(cid:19) =
a log(cid:16) V
W0(cid:17) + k
kb + aβ
4 (S0 − S)(1 − αw)
.
(5)
The warm state CO2 partial pressure has a power law dependence on the CO2 outgassing
rate. The exponent depends most strongly on a, which determines the increase in outgoing
longwave radiation with temperature. The more that increasing the warm state temperature
by a given amount increases longwave cooling, the higher the warm state CO2 must be to
maintain that temperature at a given CO2 outgassing rate. As expected, the warm state
CO2 partial pressure is exponentially lower if the stellar flux is higher. This effect is mediated
mainly by k, the parameter that determines the increase in weathering with temperature. If
k is higher, then a given warming from an increase in stellar flux causes the weathering to
increase more, and draws down the CO2 more.
Using Equation (4), we can solve for the coldest possible warm state by setting Tw = Ti.
This is a very important condition because habitability would be lost if anything were to
cool the climate when it is in the coldest possible warm state. We can solve for the stellar
flux at which the coldest possible warm state exists (S ∗), which we can think as the effective
outer edge of the habitable zone, as follows
S ∗ = S0 −(cid:18) 4
1 − αw(cid:19)(cid:18) b
β
log(cid:18) V
W0(cid:19) + k(T0 − Ti)(cid:18) b
β
+
a
k(cid:19)(cid:19) .
(6)
-- 7 --
A lower value of S ∗ means that the stellar flux must be decreased more to reach the coldest
possible warm state, which means that the warm climate state exists in more of the habitable
zone. The ideal situation for habitability is when S ∗ is decreased enough that it is smaller
than the outer edge of the habitable zone, the point at which some other process, such
as Rayleigh scattering, prevents CO2 from warming the planetary surface.
If this is the
case then the silicate-weathering feedback can keep a planet habitable throughout the entire
habitable zone.
The first thing we can note from Equation (6) is that if the albedo of the warm state is
smaller, the effective outer edge of the habitable zone will be further out. Since absorption of
stellar flux by atmospheric water vapor should be larger for planets orbiting smaller, redder
stars (Kasting et al. 1993), the effective outer edge of the habitable zone is less likely to
restrict the traditional habitable zone for M-stars than for G-stars and more likely for F-
stars, which is consistent with what has been recently found by Haqq-Misra et al. (2016).
Equation (6) also tells us that the effective outer edge of the habitable zone depends only
logarithmically on the CO2 outgassing rate. This is important because it tells us that large
changes in the CO2 outgassing rate are necessary to significantly change the effective outer
edge. For example, increasing the CO2 outgassing rate by a factor of ten only decreases the
effective outer edge from 794 W m−2 to 530 W m−2, for our standard parameters (Figure 1).
Changing the weathering parameters can have a much larger effect, which we can investigate
by doing a sensitivity analysis in which we vary them. For example, either increasing the
weathering-temperature rate constant (k) by a factor of two or decreasing the weathering-
CO2 power law exponent (β) by a factor of two causes a similar reduction in the effective
outer edge as increasing the CO2 outgassing by a factor of ten (Figure 1).
Changing k and changing β have qualitatively different effects on the response of S ∗
to changes in the CO2 outgassing rate (Figure 1). Decreasing β increases the slope of S ∗
amount extends the effective habitable zone further (Equation (6)). This is because the
temperature becomes more sensitive to the CO2 outgassing rate when β is smaller (Equa-
as a function of log(cid:16) V
tion (4)). Alternatively, changing k changes the offset of the S ∗ versus log(cid:16) V
W0(cid:17), which means that increasing the CO2 outgassing rate by a given
W0(cid:17) line, but
not the slope. A higher value of k means that the temperature has to change less to cause
the same change in weathering rate. This allows the stellar flux to be dropped to a lower
value before the warm state temperature reaches the temperature at which the warm state
is lost (Tw = Ti).
For reference I have plotted the estimated positions of modern Earth, Earth 3.8 Gyr
ago, and Mars 3.8 Gyr on Figure 1. The simple model used here indicates that early and
modern Earth are safely inside the effective outer edge of the habitable zone, whereas early
-- 8 --
Instellation at which the habitable climate state is lost
1400
1200
1000
800
600
]
2
−
m
W
[
)
∗
S
(
x
u
l
f
r
a
l
l
e
t
S
400
−1.0
Traditional HZ Inner Edge
β=0.25
k=0.2
(cid:0)
ble
bita
a
H
A
β=1.0
k=0.05
k=0.1
β=0.5
Traditional HZ Outer Edge
M
00.5
0.0
0.5
1.0
log base 10 o normalized CO2 outgassing (log(V/W0))
Fig. 1. -- The stellar flux at which the habitable climate state is lost (the effective outer
edge of the habitable zone) as a function of the logarithm of the normalized CO2 outgassing
(black line for standard parameters). The different curves show the behavior for different
values of the weathering-CO2 power law exponent (β, solid red and blue lines) and the
weathering-temperature rate constant (k, dashed red and blue lines), both of which are
relatively unconstrained. Red lines indicate changes to parameters that restrict the region
where the habitable climate state can exist and blue lines indicate changes to parameters
that expand this region. The traditional inner and outer edges of the habitable zone from
Kopparapu et al. (2013) are at the top and bottom of this plot. The symbol L represents
modern Earth, the "A" represents Earth 3.8 Gyr ago assuming 75% modern Earth's stellar
flux (i.e., insolation, Gough 1981) and a CO2 outgassing rate of three times modern, and the
"M" represents Mars 3.8 Gyr ago assuming a CO2 outgassing rate of 1 bar Gyr−1 (Grott et al.
2011), which would plot to the left of the minimum CO2 outgassing rate shown here. The
CO2 outgassing rate on early Earth is not well-constrained, but probably was higher than
modern (Dasgupta 2013).
Mars would have been beyond the effective outer edge of the habitable zone even if it were
within the traditional habitable zone. This is consistent with the histories of the two planets
if we assume that fluvial features on early Mars were episodic (Wordsworth et al. 2013;
Halevy & Head III 2014; Kite et al. 2015). That said, it should be understood that specific
-- 9 --
conclusions like these depend on details of weathering parameterizations, and the purpose
of this paper is to expose the qualitative effects of weathering, rather than to try to answer
detailed questions.
0.30
0.25
0.20
0.15
0.10
0.05
]
1
−
C
[
)
k
(
t
n
a
t
s
n
o
c
e
t
a
r
e
r
u
t
a
r
e
p
m
e
t
-
g
n
i
r
e
h
t
a
e
W
0.00
0.0
Threshold for a safe habitable zone
H
a
b
i
t
a
b
l
e
Z
o
n
e
S
a
f
V=0.1*W0
e
P ara m eter C h a n g e s to S afety
Default Values
V=W0
V=10*W0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
Weathering-CO2 power-law exponent (β)
Fig.
2. -- Lines of the weathering-temperature rate constant (k) as a function of the
weathering-CO2 power law exponent (β) at which the effective outer edge of the habit-
able zone equals the traditional outer edge. These lines are plotted for three different values
of the CO2 outgassing rate (0.1, 1, and 10 times modern Earth's value).
If the k and β
combination is to the upper left of this plot relative to a given line, then CO2 outgassing
does not limit habitability and the full habitable zone is saved.
Another way we can think about this is to consider the set of weathering parameters
that would make the effective outer edge of the habitable zone correspond to the traditional
outer edge of the habitable zone. If we denote the traditional outer edge of the habitable
zone as S ∗
out, we can rewrite Equation (6) as
k = −
log(cid:16) V
W0(cid:17)
T0 − Ti
+
β
b (cid:18) 1
4(S0 − S ∗
out)(1 − αw)
T0 − Ti
− a(cid:19) .
(7)
Equation (7) represents a series of lines of k as a function of β for different values of the CO2
outgassing rate. If k is larger or β is smaller than the line for a particular CO2 outgassing
rate, then the effective habitable zone is just as large as the traditional habitable zone for this
-- 10 --
CO2 outgassing rate. Figure 2 shows three such curves for different CO2 outgassing rates.
For the default values of k and β, some of the habitable zone would be lost even if the CO2
outgassing rate were ten times higher than modern Earth's ( V ∗
= 16.6 would be required
W0
for the effective outer edge to equal the traditional outer edge). However, relatively small
changes to k, β, or some combination of the two can save the habitable zone even for modern
Earth's CO2 outgassing rate (Figure 2). The parameters k and β are uncertain and may
vary between planets, but the expressions in this section could be used for a probabilistic
estimate of planetary habitability if appropriate priors are used for k and β.
3. What happens when the habitable climate state ceases to exist
If the stellar flux or CO2 outgassing rate is lowered enough that the warm climate state
no longer exists, the planet enters a Snowball state and the albedo increases according to
Equation (2).
It is an open question how weathering would behave in a Snowball state.
Menou (2015) assumed that weathering would go to zero due to lack of rain (liquid water),
and we will consider this case in section 4. The possibility remains, however, that weathering
could continue to occur under wet-based ice sheets or at the seafloor (Le Hir et al. 2008).
For illustrative purposes, we will continue to use Equation (3) to characterize weathering in
the Snowball state, but it should be understood that either subglacial or seafloor weathering
could lead to different parameterizations. As we will see, however, the important point
is that if some CO2-dependent weathering can occur in the Snowball state, and it causes
weathering to increase enough to balance CO2 outgassing before the Snowball deglaciates,
then it is possible for this state to be a stable solution of the system.
We can solve for the temperature tendency nullcline (the line on which dT
dt =0 in Equa-
tion (1)), which results in
log(cid:18) P
P0(cid:19) =( 1
b(cid:0)a(T − T0) + 1
b(cid:0)a(T − T0) + 1
Equation (8) describes two lines of log(cid:16) P
1
4 S0(1 − αw) − 1
4 S0(1 − αw) − 1
4 S(1 − αw)(cid:1) , T ≥ Ti
4 S(1 − αc)(cid:1) , T < Ti.
(8)
b .
The colder solution has a larger vertical offset. This is because more CO2 would be required
to keep the cold state at a given temperature than the warm state because the albedo is
higher in the cold state (although they can never actually exist at the same temperature).
Similarly, we solve for the CO2 partial pressure tendency nullcline (the line on which dP
dt =0
in Equation (3)) to find
P0(cid:17) as a function of T − T0, each with a slope of a
log(cid:18) P
P0(cid:19) =
1
β
log(cid:18) V
W0(cid:19) −
k
β
(T − T0).
(9)
-- 11 --
Equation (9) describes a line of log(cid:16) P
P0(cid:17) as a function of T − T0 with a slope of − k
β . Since
the slope is negative, there will always be at least one intersection of the two nullclines,
which will be a steady-state of the system. As the CO2 outgassing rate (V ) increases, the
intercept of Equation (9) is increased and the solution becomes warmer. For high values of
V only the warm state is a steady state, and for low values only the cold state is a steady
state. For intermediate values of V it is possible to have both states to be a steady state of
the system.
We already solved for the warm state temperature and CO2 in Equations (4) and (5).
We can now solve for the cold state temperature (Tc) and CO2 (Pc) using the cold branch
of Equation (8)
b log(cid:16) V
a log(cid:16) V
W0(cid:17) + β
W0(cid:17) + k
kb + aβ
kb + aβ
Tc − T0 =
log(cid:18) Pc
P0(cid:19) =
4 S(1 − αc) − β
4 S0(1 − αw)
4 S0(1 − αw) − k
4 S(1 − αc)
,
.
(10)
(11)
The final point of interest is to determine the stability of the fixed points described by
Equations (4), (5), (10), and (11). We can do this by evaluating the Jacobian (J) of the
system at the fixed points
J =(cid:20) ∂
∂T
∂
∂T
dT
dt
dP
dt
∂
∂P
∂
∂P
dT
dt
dP
dt(cid:21) ="
− a
C
P0(cid:17)β
−kW0ek(T −T0)(cid:16) P
− W0
P0
βek(T −T0)(cid:16) P
b
CP
P0(cid:17)β−1# .
(12)
For both the warm and cold states, the trace of the Jacobian (τ ) is less than zero and its
determinant (∆) is greater than zero. This means that the warm and cold states are always
attracting (Strogatz 1994). τ 2 − 4∆ is generally positive, which means that the fixed points
will usually be stable nodes, although it is possible for them to be stable spirals for some
parameter combinations.
I have plotted what we have learned about the nullclines and their intersections for
a representative set of parameters in Figure 3. This plot shows how intersections of the
nullclines lead to steady states, and how at least one intersection will always occur since
the CO2 partial pressure tendency nullcline has a negative slope and the two temperature
tendency nullclines have a positive slope. Note that although the simplicity of the model we
are considering constrains the nullclines to be linear, we would expect similar, but poten-
tially nonlinear, behavior from a more complicated model. For example, if we used a more
sophisticated radiative transfer model for the climate calculations (e.g., Kopparapu et al.
-- 12 --
Nullclines and fixed points of the system
=0
dP
dt
V=20*W0
V=W0
=0
dT
dt
Cold State
Warm State
V=0.05*W0
−40
−30
−20
−10
0
10
20
30
40
Temperature [ ◦ C]
10
8
6
4
2
0
−2
−4
−6
−50
e
r
u
s
s
e
r
p
l
a
i
t
r
a
p
2
O
C
d
e
z
i
l
a
m
r
o
f
o
0
1
e
s
a
b
g
o
l
dt =0, solid lines) or the CO2 partial pressure tendency is zero ( dP
Fig. 3. -- This plot shows nullclines of the system, where either the temperature tendency is
zero ( dT
dt =0, dashed lines).
Three pressure partial pressure tendency nullclines are shown: for a CO2 outgassing rate of
0.05, 1, and 20 times modern Earth's. The intersections of these nullclines represent fixed
points of the system, which are all attractors. Warm climate state fixed points are plotted in
red and cold climate state fixed points are plotted in blue. For higher CO2 outgassing rates
only the warm state exists, for lower rates only the cold state exists, and at intermediate
rates both the warm and cold climate states exist. The stellar flux is 80% of modern Earth's
in this figure.
2013) it would lead to curvature in the temperature tendency nullclines, but no change in
the topology of the system.
Because we have assumed a discontinous albedo transition (Equation (2)) stable fixed
points appear and disappear in isolation in Figure 3. If we had instead assumed a smoothed
albedo transition, for example smoothed with a hyperbolic tangent function, the two temper-
ature tendency nullclines would smoothly join together. In this case, there would always be
an unstable saddle fixed point between the two attracting fixed points representing the warm
and Snowball climate states when both attracting states exist at the same CO2 outgassing
If the CO2 outgassing rate were changed sufficiently, the saddle fixed point would
rate.
merge with either of the attracting fixed points in a saddle node bifurcation, rather than the
-- 13 --
attracting fixed point just disappearing as it does in the discontinous albedo system.
e
r
u
s
s
e
r
p
l
a
i
t
r
a
p
2
O
C
d
e
z
i
l
a
m
r
o
n
f
o
0
1
e
s
a
b
g
o
l
Climate cycles for zero Snowball weathering
=0
dT
dt
ery fast
slow
fast
ery fast
4
3
2
1
0
−1
−60
−40
=0
dP
dt
−20
Temperature [ ◦ C]
0
20
40
Fig. 4. -- Plot of the climate cycles that occur when both (1) the CO2 outgassing rate is
too low to achieve a warm climate steady state and (2) the weathering rate is set to zero
when the temperature is less than the Snowball transition temperature (T < Ti). This plot
is similar to Figure 3, but the CO2 partial pressure nullcline ends for T < Ti because there
is no way for the time derivative of the CO2 partial pressure to be zero if the weathering
rate is zero. There is no steady state and instead planetary climate experiences a limit cycle
with four stages. Most of the time is spent with the planet in the Snowball state, where it
warms very slowly as a result of CO2 outgassing.
4. Climate cycles when the Snowball weathering is set to zero
Alternatively, we can consider the situation where the weathering rate is smaller than
the CO2 outgassing rate for temperatures less than the temperature at which the planet
transitions between the two climate states (T < Ti). For simplicity, I will set the weathering
rate to zero for T < Ti, following Menou (2015). In this case no Snowball steady state is
possible because there is no weathering term to balance CO2 outgassing when the planet is
experiencing a Snowball (Equation (3)). Instead CO2 simply accumulates in the Snowball
state, warming it, until the temperature Ti is reached. This causes the albedo to decrease
-- 14 --
(physically the ice melts) and the planet abruptly jumps into the warm climate state. If we
assume that the CO2 outgassing rate is low enough that no warm climate steady state exists,
then the warm climate leads to the rapid removal of CO2 by weathering until Ti is again
reached, then the climate abruptly jumps into the Snowball state, and the cycle repeats.
Figure 4 shows a diagram of this cycle in phase space and Figure 5 shows timeseries of CO2
partial pressure and temperature through the cycle.
The two transitions between the Snowball and warm states occur on the timescale of
relaxation back to the temperature tendency nullcline. This timescale is given by C
a ≈ 3
years, which is essentially instantaneous for our purposes (the system is extremely stiff). This
allows us to make the approximation that as the CO2 changes in either the warm or Snowball
state, the climate exists along the temperature tendency nullcline ( dT
dt =0 in Equation (1)).
The CO2 partial pressure in the warm state ( Pw) as a function of the temperature in the
warm state ( Tw) is therefore
b log Pw
P0! = a( Tw − T0) +
S0
4
(1 − αw) −
S
4
(1 − αw),
(13)
and the CO2 partial pressure in the cold state ( Pc) as a function of the temperature in the
cold state ( Tc) is
b log Pc
P0! = a( Tc − T0) +
S0
4
(1 − αw) −
S
4
(1 − αc).
(14)
I have used a tilde for the CO2 partial pressure and temperature variables in these equations
because they are not true solutions of the system, since weathering never balances CO2
outgassing during the cycles.
dP
dt
is constant during the Snowball phase if the weathering rate is zero (equal to V ), so
it is easy to calculate the time spent in the Snowball phase (τc) as
τc =
Pc(Ti) − Pw(Ti)
dP
dt
≈
Pc(Ti)
dP
dt
=
P0
V
e
1
b (a(Ti −T0)+
S0
4 (1−αw)−
S
4 (1−αc)),
(15)
where we have used the fact that in general Pc(Ti) ≫ Pw(Ti). Similarly, if we assume that
silicate weathering is limited by the supply of silicate cations from erosion (Mills et al. 2011;
Foley 2015), then we can approximate the weathering rate as a constant during the warm
phase. Using similar logic as we used to get Equation (15), we arrive at a first estimate for
the time spent in the warm state (τw1) of
τw1 =
Pc(Ti) − Pw(Ti)
dP
dt
≈
Pc(Ti)
φW0 − V
≈
Pc(Ti)
φW0
= γτc,
(16)
-- 15 --
where φ is the factor by which the supply-limited maximum weathering rate exceeds modern
Earth's weathering rate (∼2.5 is a good guess for an Earth-like planet, Mills et al. 2011) and
γ is the fractional reduction in warm state relative to cold state lifetime due to the fact that
the weathering is higher in the warm state. Using W0=20V , as in Figure 4, we get γ= 1
50.
This would indicate that a negligible amount of the time in the cycle is spent in the warm
state relative to the cold state.
It is more difficult to calculate the time spent in the warm climate state if we let dP
dt
vary. Substituting into Equation (3) we find the following differential equation for the CO2
partial pressure in the warm state ( Pw)
d Pw
dt
P0!β
= V − W0ek( Tw −T0) Pw
= V − W0e
k
4a (S−S0)(1−αw ) Pw
P0!β+ kb
a
.
(17)
Equation (17) has a simple analytical solution for β + kb
a =1, which happens to be the case for
our default parameters (Table 1). I will use this limit to illustrate the behavior of the warm
state CO2 drawdown. The initial condition is Pw(t = 0) = Pc(Ti), so that Equation (17) is
solved by
Pw(t) = ( Pc(Ti) −
V
W0
P0e
k
4a (S0−S)(1−αw))e−
W0
P0
e
k
4a (S−S0)(1−αw )t +
V
W0
P0e
k
4a (S0−S)(1−αw).
(18)
We are seeking the time, τw, such that Pw(t = τw) = Pw(Ti), so plugging into Equation (18)
we find
τw2 =
P0
W0
e
k
4a (S0−S)(1−αw) log Pc(Ti) − V
W0
Pw(Ti) − V
W0
k
P0e
P0e
4a (S0−S)(1−αw)
4a (S0−S)(1−αw)! .
k
(19)
In general τc ≫ τw1 > τw2, as we would expect because the temperatures are high and the
weathering is fast in the warm state (leading to a low τw2) and we limit the weathering rate
in our other warm state timescale estimate (τw1). For example, for the parameters used
in Figure (4), τc=250 Myr (slow), the two estimates for the time spent in the warm state
are τw1=5 Myr and τw2=0.5 Myr (fast), and the time to transition between the warm and
Snowball states ( C
a ) is about 3 years (very fast).
Since the other components of the cycle take are short, τc (Equation (15)) yields a
good approximation of the period of the total cycle, which we will call τ . We can drop the
constants associated with the reference state in Equation (15) as follows to think about the
variable dependencies of τ
τ ∝
a
b Ti−
S
4b (1−αc).
e
(20)
1
V
As one might expect, the period of the cycles is inversely proportional to the CO2 outgassing
rate. The more interesting aspect of Equation (20) is that it gives us a functional form
-- 16 --
Climate cycles fo ze o Snowball weathe ing
200
400
600
800
1000
200
400
600
Time [My ]
800
1000
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
]
a
b
[
e
u
s
s
e
p
l
a
i
t
a
p
2
O
C
0.0
0
40
20
0
−20
−40
−60
0
]
C
◦
[
e
u
t
a
e
p
m
e
T
Fig. 5. -- Timeseries of CO2 partial pressure and temperature for the limit cycle depicted in
phase space in Figure 4. Most of the limit cycle is spent in the Snowball state, during which
the CO2 increases linearly in the atmosphere. I have made this plot based on the analytical
expressions in section 4: exploiting the constant CO2 accumulation in the Snowball state,
a warm state CO2 given by Equation (18), temperature jumps at constant CO2 for the
transitions between states, and using Equations (13) and (14) to find the temperatures from
the CO2 values.
for the dependence of the period of the cycle on the stellar flux and albedo of the cold
state. If we start in the warm state and decrease the stellar flux (say by moving the planet
-- 17 --
away from the star), a limit cycle will suddenly appear when the threshold for existence of
the warm state is crossed (Equation (6)). If we continue to decrease the stellar flux, the
period of this cycle will grow exponentially as the stellar flux is decreased. The exponential
functional dependence ultimately derives from the logarithmic effect of CO2 on infrared
emission to space (Equation (1)). This means that the period of the cycle for planets near
the outer edge of the habitable zone will be very long, so the difference between having a
permanent Snowball state as in section 3 and having very slow cycles through the Snowball
and warm states as described in this section will be marginal. Similarly, the timescales
will be exponentially shorter as we decrease the Snowball albedo. Since ice and snow have
a much lower albedo for an M-star spectrum (Joshi & Haberle 2012; Shields et al. 2013),
this implies that the climate cycles for planets in M-star systems would have much shorter
periods. Finally, increasing b decreases the effect of changing the stellar flux and Snowball
albedo. It is important to note that since we have set the weathering rate to zero in the
Snowball, which is the part of the cycle that determines its period, none of the uncertain
weathering parameters affect the period of the cycle.
The dependence of τ on Ti (Equation (20) can significantly affect the comparison of
different models. For example, I implemented a smoothed version of the albedo transition
(Equation (2)) and integrated the system numerically. I found that the period of the cycle was
strongly dependent on the temperature smoothing of the albedo parameterization because
this affected the effective temperature at which the transition from the Snowball to the warm
climate occurred. Using a
b = 0.2, a decrease in Ti by about 10 K leads to a decrease in τ
by about an order of magnitude. This sensitivity will affect the comparison of cycle periods
between models, however, the dependencies shown in Equation (20) should hold within a
given model.
We can use Equation (20) to understand simulations in more complex models. For
example, Menou (2015) performed six different simulations of climate cycles for planets at
different orbital distances (corresponding to different stellar fluxes), CO2 outgassing rates,
and values of β. Two of these simulations have a CO2 outgassing rate three times higher
than the others, and, according to Equation (20), I have adjusted their period by multiplying
it by three. I have plotted the logarithm of the adjusted period as a function of stellar flux in
Figure 6. I have also plotted the time spent in the Snowball state for each simulation, which
we can calculate because Menou (2015) gives the percentage of the cycle spent in the warm
state. The first thing to note is that the logarithm of the period is fairly linear in stellar flux,
consistent with Equation (20). If we assume that the Snowball albedo is 0.7, the slope of the
line corresponds to a value of b of about 27 W m−2, which is the right order of magnitude.
Two of the simulations have periods 10 -- 20% longer than expected. These simulations have
a lower value of β, and therefore spend a higher fraction of their cycle in the warm state
-- 18 --
(Menou 2015). We should also note that even the time spent in the Snowball state does
not fall exactly on a straight line, and it is not a single-valued function of stellar flux. The
reasons for this are: (1) Menou (2015) uses a much more complex radiation model in which
infrared emission to space is not simply a linear function of the logarithm of CO2, (2) Menou
(2015) calculates the Snowball albedo including the effect of the amount of CO2, so it is not
a constant, and (3) the weathering parameterization that Menou (2015) uses allows some
weathering for temperatures slightly below the Snowball deglaciation temperature threshold,
which can delay the deglaciation and causes different results for different values of β. Despite
these differences, Equation (20) does an excellent job of describing the qualitative behavior
of period of the cycles from the simulations of Menou (2015).
2.8
2.6
2.4
2.2
2.0
]
r
y
M
[
e
l
c
y
c
f
o
d
o
i
r
e
p
f
o
0
1
e
s
a
b
g
o
l
1.8
400
500
600
700
900
Stellar Flux [W~m−2]
800
1000
1100
1200
Fig. 6. -- Logarithm of the period of the climate cycles from Menou (2015) (red circles) and
time in the Snowball state (black circles) as a function of stellar flux. Lines of best fit, with
corresponding colors, are also shown. Two of the simulations from Menou (2015) have a CO2
outgassing rate three times higher than the other simulations, and I adjusted the period of
these simulations by multiplying it by three.
-- 19 --
5. Discussion
A major advance of this paper is to derive an explicit formula for the effective outer
edge of the habitable zone (Equation 6). Although the weathering parameters in this formula
are uncertain, it could be incorporated into future probabilistic estimates of habitability of
discovered exoplanets, with appropriate prior distributions placed on weathering parameters.
Once a planet is beyond this limit, I have found that it will either experience a permanent
Snowball state or long cycles between a Snowball and warm climate, depending on whether
weathering goes completely to zero during the Snowball or note. Either way the habitability
of the planet would be greatly reduced. Also, the fact that the Snowball episodes Earth has
experienced did end does not imply that the weathering was zero during them (Le Hir et al.
2008), since the stellar flux was relatively high during these episodes.
Kopparapu et al. (2014) found that the outer edge of the traditional habitable zone
has only a small dependence on planet size, but the effective outer edge of the habitable
zone could strongly depend on planet size. A larger planet will tend to have a higher rate of
volcanism, and presumably CO2 outgassing, yet it will also have a larger overburden pressure
for a given volatile inventory (Kite et al. 2009). These competing effects will determine how
planetary size affects CO2 outgassing rate, and consequently susceptibility to loss of the
warm climate state inside the traditional habitable zone. Moreover, the CO2 outgassing
rate should decrease strongly with time (Kite et al. 2009), which indicates that planets near
the outer edge of the traditional habitable zone are more likely to actually be habitable in
younger systems.
The albedo and thermal phase curve of an Earth-like planet could be interrogated to
determine whether it was in a Snowball or warm climate state (Cowan et al. 2012). This
might be possible for a planet near the outer edge of the habitable zone with the James Webb
Space Telescope (Yang et al. 2013; Koll & Abbot 2015), and would certainly be possible with
a future mission such as the High Definition Space Telescope (Dalcanton et al. 2015). New
geochronological data (Condon et al. 2016) suggest that the Sturtian and Marinoan Snowball
Earth episodes had a combined duration of about 80 Myr, which is about 10% of the time
since they occurred, implying that a roughly 10% Snowball duty cycle could be realistic for
an Earth-like planet. Planets near the outer edge of the traditional habitable zone that have
a stable warm state but are perturbed away from it and into a Snowball could take longer
to warm up via CO2 outgassing and will therefore spend a somewhat higher percentage of
their time as a Snowball.
If these planets are anything like Earth, however, they should
still spend the vast majority of their time in the warm climate state. Planets outside the
effective outer edge of the habitable zone, on the other hand, should spend most of their
time as a Snowball. Therefore a measurement of the fraction of Earth-like planets in a warm
-- 20 --
state as a function of position in the habitable zone would tell us whether CO2-outgassing
limitations on the habitable zone are important. A large increase in the average number
of Earth-like planets in a Snowball state near the outer edge of the habitable zone would
indicate that CO2-outgassing limitations are important on average. In this way astronomical
measurements could increase our understanding of weathering. Similarly, continued study of
river catchments, paleoclimate, and laboratory weathering analogs should help improve our
understanding of weathering, and therefore the effective outer edge of the habitable zone,
which will inform the astronomical search for habitable exoplanets.
In this paper I have not included the effect of CO2 on planetary albedo. This means that
if the weathering is set to zero in a Snowball state, then the CO2 can always build up to high
enough values to cause deglaciation and lead to climate cycles if the warm state does not
exist (section 4). For Snowball states that require tens of bars of CO2 to deglaciate, increased
shortwave scattering by CO2 could prevent deglaciation from ever occurring. If this were the
case, the Snowball state could be stable even if the CO2 cycle is not balanced. If deglaciaton
does occur at a very high CO2 level, then the atmospheric albedo might be so high that the
change in surface albedo has a minimal effect on the planetary albedo (Wordsworth et al.
2011), so the warming associated with deglaciation is minimal. In this case CO2 could still
be drawn down due to high CO2 concentrations (Equation (3)), leading to a climate cycle,
but the planet would likely spend more time in the warm state.
Finally, we should note that even if the effective outer edge of the habitable zone oc-
curs at a significantly higher stellar flux than the traditional outer edge, there should still
be plenty of habitats for life in the universe.
If the habitable zone were cut in half, the
proportion of Sun-like stars hosting Earth-like potentially habitable planets would go from
∼5% to ∼2.5% (Petigura et al. 2013), which might have been considered an optimistic esti-
mate before the Kepler mission. Moreover, recent work on H2-greenhouse planets suggests
that habitable planets can exist even outside of the traditional habitable zone (Stevenson
1999; Pierrehumbert & Gaidos 2011; Wordsworth 2012; Abbot 2015). Additionally, life, and
maybe even animal life, seems to have survived the Snowball Earth episodes, and could po-
tentially survive permanent or cyclical Snowball climates near the outer edge of the habitable
zone, although such conditions would certainly be less favorable to complex life than modern
Earth. Finally, if simple life can exist in subglacial oceans on distant or unbound Earth-like
planets (Laughlin & Adams 2000; Abbot & Switzer 2011), then the Snowball planets beyond
the effective habitable zone would still be viable hosts for some sort of life.
-- 21 --
6. Conclusions
The main conclusions of this paper are:
1. The stellar flux at the effective outer edge of the habitable zone can be approximated
by the following formula:
S ∗ = S0 −(cid:18) 4
1 − αw(cid:19)(cid:18) b
β
log(cid:18) V
W0(cid:19) + k(T0 − Ti)(
b
β
+
a
k
)(cid:19) .
Larger values of k, the weathering-temperature rate constant, linearly decrease the
stellar flux of the effective outer edge of the habitable zone, moving it farther from
the star and providing more habitable space in the system. Smaller values of β, the
weathering-CO2 power law exponent, directly decrease the stellar flux of the effective
outer edge of the habitable zone and also leverage the effect of increases in the CO2
outgassing rate. If k is increased by about a factor of two or β is decreased by a factor
of two, or some smaller combination of the two, then the effective outer edge of the
habitable zone equals the traditional outer edge of the habitable zone even for modern
Earth's CO2 outgassing rate, and none of the habitable zone would be lost. These
changes are within the uncertainty in the values of k and β. This equation also tells us
that M-star planets should tend to have less of a reduction in their habitable zone due
to limited CO2 outgassing, since αw, the warm state albedo, will tend to be smaller
for M-star planets (making S ∗ smaller).
The formula for S ∗ could be incorporated into probabilistic estimates of whether a dis-
covered exoplanet is habitable, using appropriate priors on k and β. It could similarly
be used to estimate the fraction of stars that host an Earth-like planet given statistics
of exoplanet occurrences. Both of these uses would aid in the planning of telescopes
that would observe Earth-like planets and search for biosignatures.
2. Beyond the effective outer edge of the habitable zone (but inside the traditional outer
edge) cycles between a Snowball and warm climate are possible if weathering is weak
enough that the CO2 needed to deglaciate a Snowball is reached before weathering can
balance CO2 outgassing (for example if the weathering rate is simply set to zero in
a Snowball climate). If weathering occurs either subglacially or at the seafloor, it is
possible to have a stable, attracting Snowball climate state.
3. If climate cycles between a Snowball and warm state occur, then the period of these
cycles scales as
τ ∝
1
V
a
b Ti−
S
4b (1−αc)).
e
-- 22 --
This formula comes from the time spent in the Snowball state, which dominates the to-
tal period and can be calculated by dividing the CO2 needed to deglaciate the Snowabll
by the CO2 outgassing rate (which explains why the period of the cycles is inversely
proportional to the CO2 outgassing rate). The exponential dependence on the temper-
ature at which the Snowball state deglaciates and the negative exponential dependence
on the stellar flux ultimately derive from the fact that CO2 has a logarithmic effect
on infrared emission to space and the greenhouse warming of a planet. The negative
exponential dependence on stellar flux indicates that cycles near the outer edge of the
habitable zone will have very long periods, and may be hard to distinguish from per-
manent Snowball states. The exponential dependence on the planetary albedo of the
cold state indicates that climate cycles will have a shorter period for planets orbiting
M-stars because the albedo of ice and snow is lower for an M-star spectrum. Finally, it
is important to note that none of the uncertain weathering parameters appear in this
scaling.
7. Acknowledgements
I acknowledge support from the NASA Astrobiology Institutes Virtual Planetary Labo-
ratory, which is supported by NASA under cooperative agreement NNH05ZDA001C. I thank
Navah Farahat for helping me learn python plotting routines. I thank Cael Berry, Edwin
Kite, Daniel Koll, Mary Silber, and Robin Wordsworth for reading an early draft of this
paper and providing detailed and insightful suggestions.
Abbot, D. S. 2015, Astrophysical Journal Letters, 815, L3
REFERENCES
Abbot, D. S., Cowan, N. B., & Ciesla, F. J. 2012, Astrophysical Journal, 756, 178,
doi:10.1088/0004-637X/756/2/178
Abbot, D. S., & Switzer, E. R. 2011, Astrophysical Journal, 735, L27, doi:10.1088/2041-
8205/735/2/L27
Abbot, D. S., Voigt, A., & Koll, D. 2011, Journal of Geophysical Research, 116, D18103,
doi:10.1029/2011JD015927
Berner, R. 1994, Am J Sci, 294, 56
-- 23 --
Berner, R. A. 2004, The Phranerozoic Carbon Cycle (Oxford University Press, New York,
N.Y.)
Condon, D., Macdonald, F. A., Rooney, A. D., Zhu, M., Schmitz, M. D., & Bowring, S. A.
2016, Sci. Adv., submitted
Cowan, N. B., Abbot, D. S., & Voigt, A. 2012, Astrophysical Journal, 757, 80,
doi:10.1088/0004-637X/757/1/80
Dalcanton, J., et al. 2015, arXiv preprint arXiv:1507.04779
Dasgupta, R. 2013, Rev Mineral Geochem, 75, 183
Foley, B. J. 2015, The Astrophysical Journal, 812, 36
Goldblatt, C., Robinson, T. D., Zahnle, K. J., & Crisp, D. 2013, Nature Geoscience, 6, 661
Goldblatt, C., & Watson, A. J. 2012, Philosophical Transactions Of The Royal Society A-
Mathematical Physical And Engineering Sciences, 370, 4197
Gough, D. O. 1981, Sol Phys, 74, 21
Grott, M., Morschhauser, A., Breuer, D., & Hauber, E. 2011, Earth and Planetary Science
Letters, 308, 391
Halevy, I., & Head III, J. W. 2014, Nature Geoscience, 7, 865
Haqq-Misra, J., Kopparapu, R. K., Batalha, N. E., Harman, C. E., & Kasting, J. F. 2016,
arXiv preprint arXiv:1605.07130
Hoffman, P. F., Kaufman, A. J., Halverson, G. P., & Schrag, D. P. 1998, Science, 281, 1342
Huybers, P., & Langmuir, C. 2009, Earth and Planetary Science Letters, 286, 479
Joshi, M. M., & Haberle, R. M. 2012, Astrobiology, 12, 3
Kadoya, S., & Tajika, E. 2014, The Astrophysical Journal, 790, 107
Kasting, J. F. 1988, Icarus, 74, 472
Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108
Kirschvink, J. 1992, in The Proterozoic Biosphere: A Multidisciplinary Study, ed. J. Schopf
& C. Klein (Cambridge University Press, New York), 51 -- 52
-- 24 --
Kite, E., Manga, M., & Gaidos, E. 2009, Astrophys Journal, 700, 1732, doi:10.1088/0004-
637X/700/2/1732
Kite, E. S., Howard, A. D., Lucas, A. S., Armstrong, J. C., Aharonson, O., & Lamb, M. P.
2015, Icarus, 253, 223
Koll, D. D. B., & Abbot, D. S. 2015, The Astrophysical Journal, 802, 21, doi:10.1088/0004-
637X/802/1/21
Kopparapu, R. K. 2013, Astrophysical Journal Letters, 767, L8
Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J., Kasting, J. F., Domagal-Goldman,
S., & Eymet, V. 2014, The Astrophysical Journal Letters, 787, L29
Kopparapu, R. K., et al. 2013, The Astrophysical Journal, 765, 131
Laughlin, G., & Adams, F. 2000, Icarus, 145, 614
Le Hir, G., Ramstein, G., Donnadieu, Y., & Godderis, Y. 2008, Geology, 36, 47
Leconte, J., Forget, F., Charnay, B., Wordsworth, R., & Pottier, A. 2013a, Nature, 504, 268
Leconte, J., Forget, F., Charnay, B., Wordsworth, R., Selsis, F., Millour, E., & Spiga, A.
2013b, Astronomy And Astrophysics, 554, A69
Menou, K. 2015, Earth and Planetary Science Letters, 429, 20
Mills, B., Watson, A. J., Goldblatt, C., Boyle, R., & Lenton, T. M. 2011, Nature Geosciences,
4, 861, DOI: 10.1038/NGEO1305
Nakajima, S., Hayashi, Y. Y., & Abe, Y. 1992, Journal of the Atmospheric Sciences, 49,
2256
Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Proceedings of the National Academy
of Sciences, 110, 19273
Pierrehumbert, R., & Gaidos, E. 2011, Astrophys J Lett, 734, L13, doi:10.1088/2041-
8205/734/1/L13
Pierrehumbert, R. T. 2010, Principles of Planetary Climate (Cambridge University Press)
Rose, B. E. 2015, Journal of Geophysical Research: Atmospheres, 120, 1404
Shields, A. L., Meadows, V. S., Bitz, C. M., Pierrehumbert, R. T., Joshi, M. M., & Robinson,
T. D. 2013, Astrobiology, 13, 715
-- 25 --
Stevenson, D. J. 1999, Nature, 400, 32, doi:10.1038/21811
Strogatz, S. 1994, Nonlinear dynamics and chaos (Westview Press)
Tajika, E. 2007, Earth, 59, 293
Voigt, A., & Abbot, D. S. 2012, Climate of the Past, 8, 2079, doi:10.5194/cp-8-2079-2012
Voigt, A., Abbot, D. S., Pierrehumbert, R. T., & Marotzke, J. 2011, Clim. Past, 7, 249,
doi:10.5194/cp-7-249-2011
Walker, J. C. G., Hays, P. B., & Kasting, J. F. 1981, Journal of Geophysical Research, 86,
9776
West, A. J., Galy, A., & Bickle, M. 2005, Earth and Planetary Science Letters, 235, 211
Wolf, E., & Toon, O. 2014, Geophysical Research Letters,
41,
167172, dOI:
10.1002/2013GL058376
-- . 2015, Journal of Geophysical Research, 120, 5775
Wordsworth, R. 2012, Icarus, 219, 267, doi:10.1016/j.icarus.2012.02.035
Wordsworth, R., Forget, F., Millour, E., Head, J. W., Madeleine, J. B., & Charnay, B. 2013,
Icarus, 222, 1
Wordsworth, R. D., Forget, F., Selsis, F., Millour, E., Charnay, B., & Madeleine, J.-B. 2011,
Astrophys J Lett, 733, L48, doi:10.1088/2041-8205/733/2/L48
Yang, J., Bou´e, G., Fabrycky, D. C., & Abbot, D. S. 2014, Astrophysical Journal Letters,
787, L2, doi:10.1088/2041-8205/787/1/L2
Yang, J., Cowan, N. B., & Abbot, D. S. 2013, Astrophysical Journal Letters, 771, L45,
DOI:10.1088/2041-8205/771/2/L45
Yang, J., Peltier, W., & Hu, Y. 2012a, Journal of Climate, 25, 2711, doi: 10.1175/JCLI-D-
11-00189.1
-- . 2012b, Journal of Climate, 25, 2737, doi: 10.1175/JCLI-D-11-00190.1
Yang, J., Peltier, W. R., & Hu, Y. 2012c, Climate of the Past, 8, 907, doi:10.5194/cp-8-907-
2012
This preprint was prepared with the AAS LATEX macros v5.2.
|
1606.08923 | 1 | 1606 | 2016-06-29T00:41:16 | NEOWISE Reactivation Mission Year Two: Asteroid Diameters and Albedos | [
"astro-ph.EP"
] | The Near-Earth Object Wide-Field Infrared Survey Explorer (NEOWISE) mission continues to detect, track, and characterize minor planets. We present diameters and albedos calculated from observations taken during the second year since the spacecraft was reactivated in late 2013. These include 207 near-Earth asteroids and 8,885 other asteroids. $84\%$ of the near-Earth asteroids did not have previously measured diameters and albedos by the NEOWISE mission. Comparison of sizes and albedos calculated from NEOWISE measurements with those measured by occultations, spacecraft, and radar-derived shapes shows accuracy consistent with previous NEOWISE publications. Diameters and albedos fall within $ \pm \sim20\%$ and $\pm\sim40\%$, 1-sigma, respectively, of those measured by these alternate techniques. NEOWISE continues to preferentially discover near-Earth objects which are large ($>100$ m), and have low albedos. | astro-ph.EP | astro-ph | NEOWISE Reactivation Mission Year Two: Asteroid Diameters
and Albedos
C. R. Nugent1, A. Mainzer2, J. Bauer2, R. M. Cutri1, E. A. Kramer2, T. Grav3, J.
Masiero2, S. Sonnett2, and E. L. Wright4
Received
;
accepted
1Infrared Processing and Analysis Center, California Institute of Technology, Pasadena,
CA 91125, USA
2Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109 USA
3Planetary Science Institute, Tucson, AZ
4Department of Physics and Astronomy, University of California, Los Angeles, CA 90095,
USA
6
1
0
2
n
u
J
9
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
3
2
9
8
0
.
6
0
6
1
:
v
i
X
r
a
ABSTRACT
The Near-Earth Object Wide-Field Infrared Survey Explorer (NEOWISE)
mission continues to detect, track, and characterize minor planets. We present
diameters and albedos calculated from observations taken during the second year
since the spacecraft was reactivated in late 2013. These include 207 near-Earth
asteroids and 8,885 other asteroids. 84% of the near-Earth asteroids did not have
previously measured diameters and albedos by the NEOWISE mission. Compar-
ison of sizes and albedos calculated from NEOWISE measurements with those
measured by occultations, spacecraft, and radar-derived shapes shows accuracy
consistent with previous NEOWISE publications. Diameters and albedos fall
within ± ∼ 20% and ± ∼ 40%, 1-sigma, respectively, of those measured by these
alternate techniques. NEOWISE continues to preferentially discover near-Earth
objects which are large (> 100 m), and have low albedos.
1.
Introduction
Observing asteroids at infrared wavelengths is an effective method for calculating
diameters for large numbers of asteroids. Since asteroid albedos can vary by approximately
an order of magnitude, sizes estimated from reflected visible light fluxes alone have large
uncertainties. Combining diameters calculated from infrared fluxes with visible magnitudes
yields albedo measurements. Together, diameter and albedo measurements are basic
physical characterizations that enable further investigations, including studies of asteroid
families (Masiero et al. 2013; Walsh et al. 2013; Carruba et al. 2013; Milani et al. 2014;
Masiero et al. 2015a,b) and size-frequency distributions (Zellner 1979; Gradie & Tedesco
1982; Tedesco et al. 2002; Bus & Binzel 2002; Mainzer et al. 2011b; Grav et al. 2012a, 2011;
Bauer et al. 2013).
2
We present diameters and albedos of asteroids from the second year of the NEOWISE
mission following the reactivation of the spacecraft from hibernation in late 2013. Diameters
and albedos of asteroids from the first year of the NEOWISE mission following reactivation
are given in Nugent et al. (2015). NEOWISE is a space-based infrared telescope that
obtains an image of the sky every eleven seconds simultaneously in two bands, W1 (3.4
µm) and W2 (4.6 µm). From its sun-synchronous orbit around Earth, NEOWISE observes
the entire static sky every six months. The original mission, WISE, is described in detail
in Wright et al. (2010), and the NEOWISE enhancement to the mission is described in
Mainzer et al. (2011a). After successfully completing its prime mission in 2011, the WISE
spacecraft was placed into hibernation for 32 months before being reactivated and renamed
NEOWISE in late 2013. The NEOWISE reactivation mission is described in Mainzer et al.
(2014a).
The goals of the NEOWISE mission are to discover, track, and characterize minor
planets. Images and extracted source lists from all phases of the WISE and NEOWISE
missions have been delivered to the public via the Infrared Science Archive (Cutri et al.
2012, 2015), NASA's designated archive for infrared astronomical data.
During the initial portion of the mission, NEOWISE employed four channels; 3.4, 4.6,
12, and 22 µm. The longest two wavelength channels required cooling to < 8 K using a
dual-stage solid hydrogen cryostat. Diameters and albedos for a variety of small body
populations were calculated using this fully cryogenic portion of the mission (see Table 1).
As the cryogen was depleted, the 12 and 22 µm channels became inoperative; after this, the
mission continued for several months using only its 3.4 and 4.6 µm channels. A summary
of near-Earth asteroid (NEAs) and main belt asteroid (MBA) albedos and diameters
calculated during various phases of the mission is given in Table 2. These measurements
have also been submitted to NASA's Planetary Data System. Thermal model calibration
results, including comparison of cryogenic WISE/NEOWISE-derived diameters to other
3
observations, are given in Mainzer et al. (2011c,d).
This second year of data also provides multi-epoch observational data of uniform quality
that can be used to better constrain the sizes, shapes, rotation state and thermophysical
properties of the 9,092 asteroids in the reactivation Year 2 sample.
We present preliminary diameters and albedos calculated from NEOWISE Year 2
Reactivation mission observations, which spanned 13 December 2014 to 13 December 2015.
Diameters and albedos calculated from NEOWISE Year 2 Reactivation mission observations
will be submitted to the Planetary Data System.
2. Discoveries and follow-up
NEOWISE discovered 198 near-Earth asteroids and comets during Years 1 and 2 of
the Reactivation mission. In addition to observing 175 NEAs that had not had diameters
measured previously from NEOWISE data, the Year 2 Reactivation mission obtained
thermal infrared observations at additional epochs for 32 NEAs. NEOWISE typically
observes asteroids ∼ 10 − 12 times over ∼ 1 − 1.5 days, and requires a minimum of 5
detections of a discovery candidate for submission to the Minor Planet Center (MPC).
NEOWISE observes with a fixed observing cadence, and additional follow up
observations are usually necessary to confirm that new minor planet candidates have
been discovered. Since NEOWISE cannot perform targeted follow up on its own, these
observations must be made by ground-based observers. Given that near-Earth objects
(NEOs) are of a population of special interest, NEOWISE candidate NEOs are listed on the
MPC Near-Earth Object Confirmation Page (NEOCP) to facilitate follow up. NEOWISE
regularly relies on many ground-based observers for follow up, including Spacewatch,
observers at the Institute for Astronomy at the University of Hawaii, the Las Cumbres
Observatory Global telescope Network (LCOGT), the Magdalena Ridge Observatory, the
4
Table 1. Diameters and albedos for various small body populations, calculated from fully
cryogenic (3.4, 4.6, 12 and 22 µm bands) NEOWISE mission data.
Population
Associated reference
Near-Earth asteroids Mainzer et al. (2011b), Mainzer et al. (2014b)
Main belt asteroids Masiero et al. (2011)
Active main belt objects Bauer et al. (2012b)
Trojans Grav et al. (2012b)
Hildas Grav et al. (2012a)
Irregular satellites Grav et al. (2015)
Centaurs Bauer et al. (2013)
Comets Bauer et al. (2011, 2012a, 2015),
Table 2. Previously published papers containing NEA and MBA diameters and albedos
calculated from NEOWISE mission data.
Mission Phase
Detection bands (µm)
Reference for NEAs
Reference for MBAs
Fully cryogenic
3.4, 4.6, 12, 22
Mainzer et al. (2011b), Masiero et al. (2011),
Mainzer et al. (2014b) Masiero et al. (2014)
3-Band and Post cryogenic
3.4, 4.6, (some 12)
Mainzer et al. (2012a) Masiero et al. (2012)
Year 1 of reactivation
3.4, 4.6
Nugent et al. (2015)
Nugent et al. (2015)
5
Mt. John Observatory, and a number of amateur observers across the globe to coordinate
follow up of particular objects. The NEOWISE team was granted eight hours each semester
of Target of Opportunity observing time on Gemini Observatory's Gemini South telescope
(Hook et al. 2004) as well as time on the Blanco 4m/Dark Energy Camera (DECam;
Flaugher et al. 2015), and was granted Co-I status on the LCOGT NEO follow up program.
Access to these facilities is vital for following up of discoveries deep in the Southern
Hemisphere.
Figure 1 is a histogram of the observatories and campaigns that contributed the
majority of follow observations occurring immediately after NEOWISE reported candidate
observations, including Spacewatch (McMillan 2007), with over 100 follow-up observations,
the Las Cumbres Observatory Global Telescope Network (Brown et al. 2013), and the
Catalina Sky Survey (Christensen et al. 2015). Observers on Mauna Kea using the
University of Hawaii 2.2m telescope and the Canada-France-Hawaii Telescope's Megacam
imager repeatedly obtained follow-up of objects under challenging observing conditions
(e.g. Tholen et al. 2014). The Mt John University Observatory (observatory code 474)
obtained valuable observations of 2015 OA22. NEOWISE discovered this object at −70◦
declination, and ephemerides showed it was moving further South. The Mt John University
Observatory was able to track the object to −78◦ declination, confirming the discovery.
NEOWISE submitted several candidate objects to the MPC that were not placed on
the NEOCP based on an initial orbit determination that indicated they were not NEOs.
These did not receive targeted follow-up. Fifty-seven non-NEO NEOWISE discoveries
made during the Reactivation mission do not have associated orbits. Additionally, there
were eight objects (main belt asteroids, Hungarias, and Mars-crossers) detected solely by
NEOWISE that were given provisional designations by the MPC; all of these objects have
poorly determined orbits. Without well-determined orbits, distance at observing time
cannot be computed accurately, and therefore diameters were not determined for these
6
Fig. 1.- NEOWISE relies on ground-based observatories to perform targeted follow up of
candidate NEO discoveries. This histogram shows numbers of follow-up observations taken
by different groups (listed by their MPC observatory codes) for NEOWISE discoveries during
Year 1 and Year 2 of the Reactivation mission.
objects.
7
2.1. Comets
The NEOWISE Reactivation mission has detected over 100 comets, including eight
discoveries (four of which were made after the end of the second year of observations).
The NEOWISE spacecraft is sensitive to the presence of coma dust, as well as the CO-line
(4.67 µm) and CO2-line (4.23 µm) emission from comet comae from gas species which are
obscured or completely blocked by Earth's atmosphere (Figure 2). Analysis of the excess
emission at 3.4 µm by Bauer et al. (2015) provided CO+CO2 production rates and limits
of the first four comets discovered by the NEOWISE Reactivation mission.
The infrared wavelengths provide a thermal emission and reflected light dust signal
that can characterize a unique regime of dust particle sizes through analysis of the dust
coma morphologies (Kramer et al. 2015). The NEOWISE multi-epoch observations of
many of the comets detected so far provide characterization of long-term cometary behavior
regarding these aspects of dust and gas emission. The gas and dust properties of the
Reactivation Year 1 and 2 survey comet sample will be described in a later work.
3. Methods
3.1. Extraction of Detections
The methodology for extracting detections of minor planets from the NEOWISE source
lists, as well as methods of diameter and albedo computation follows the description in
Nugent et al. (2015), with the one exception described in Section 3.2. As was done in that
work, the NEOWISE source lists were searched using the positions and times reported for
each minor planet in the MPC's archival files NumObs.txt and UnnObs.txt1. NEOWISE
reports detections to the MPC three times a week. By querying the MPC archive after
1http://www.minorplanetcenter.net/iau/ECS/MPCAT-OBS/MPCAT-OBS.html
8
Fig. 2.- Comet C/2013 US10 (Catalina) as seen by NEOWISE on August 28, 2015. The
2-color image maps the 3.4 µm band to the cyan, and the 4.6 µm band to the red. The image
is a quarter-degree on a side and is oriented approximately with North down and East is to
the left. The red appearance suggests the comet may have significant CO or CO2 emission.
the conclusion of Year 2 operations, we restrict our analysis to those detections of minor
planets that were reported to and confirmed by the MPC.
These detections were converted into IRSA Catalog Query Engine format2, and were
used to query the NEOWISE-R Single Exposure Level 1b (L1b) Source Table, which is
served by the NASA/IPAC Infrared Science Archive (IRSA). The NEOWISE Reactivation
2http://irsa.ipac.caltech.edu/applications/Gator/GatorAid/irsa/QuickGuidetoGator.htm
9
data are described in detail in the NEOWISE Reactivation Explanatory Supplement (Cutri
et al. 2015), which was updated in March 2016 to include single-exposure images and
extracted source products from Reactivation Year 2. The Single Exposure (L1b) Source
table was queried to find sources within 2 arcseconds of the reported position in the MPC
files. For this query, detection time is constrained to be within two seconds of the reported
time. The resulting table is a list of all sources corresponding to reported MPC detections
from single exposures, with associated MPC designations for each detection.
Several steps are taken to prevent confusion of small body detections with fixed
background sources such as stars and galaxies. We reference the WISE All-Sky Source
Catalog, which is derived from a co-add of multiple exposures, covering the sky. This is
a significantly deeper image than the individual L1b images, and pixel outlier rejection
suppresses moving solar system objects. Therefore, it is useful for identifying fixed sources
in the L1b images. The WISE Moving Object Pipeline System (WMOPS), which identifies
moving objects in the NEOWISE images, compares single-exposure detections to reference
images before any detections are submitted to the MPC. However, as an additional
precaution, we also compare the single-exposure detection list to the All-Sky Catalog. Any
single-exposure detections found to be within 6.5 arcseconds (the size of the 3.4 and 4.6 µm
NEOWISE point-spread function) of a WISE All-Sky Source Catalog source with SNR ≥ 3
were removed.
The resulting asteroid detection table was then stripped of measurements with
associated poor quality flags. Each NEOWISE detection is graded for quality, as described
in Cutri et al. (2015). Detections with "ph qual" values of "A", "B", or "C" were accepted,
this photometric quality grade ensures that the source was detected in the band with a flux
signal-to-noise ratio < 2. Additionally, detections must have "cc f lags" values of "0" or
"p", indicating that either the source was unaffected by known artifacts ("0"), or perhaps is
impacted by a latent image left by a bright source ("p"). The value of "p" is conservative;
10
it indicates the source is likely unaffected by a latent image, but possibly may be slightly
contaminated. Finally, only frames graded "qual f rame'="10" or highest quality by the
quality assurance process were used.
The WISE Science Data System pipeline profile-fitting magnitudes are used for each
band (Cutri et al. 2015). A minimum of three detections with measurement uncertainties
σmag ≤ 0.25 mag were required for thermal fits. Saturated detections, with a W1 magnitude
≤ 8.0 or a W2 magnitude ≤ 7.0, were discarded. The photometric measurements used for
each asteroid are listed in Table 3.
3.2. H and G values
For each diameter, a corresponding albedo is also calculated, using an absolute visual
magnitude H and IAU phase slope parameter G (Bowell et al. 1989). Therefore, the
accuracy of albedos calculated from diameter measurements depends on the accuracy H
and G values. The MPC provides H and G values as part of its catalog service; however,
the default catalog values may be affected by various systematic effects (Veres et al. 2015;
Williams 2012). Known issues include values calculated from observations submitted with
uncertain photometric calibrations and a bias towards discovering asteroids when their
longest axis faces Earth (Jedicke et al. 2002).
Corrected or newly-derived H and G values have been published by Warner et al.
(2009); Pravec et al. (2012); Williams (2012) and Veres et al. (2015). The largest of these H
and G datasets is from Williams (2012) with ∼ 337000 numbered asteroids, and Veres et al.
(2015) with ∼ 250000 objects observed by Pan-STARRS PS1. The Williams (2012) dataset
is slated to be incorporated into the MPC catalog (G. Williams, personal communication,
May 2nd 2016). For this reason, and because it is more extensive, we used these corrected
H and G values in this work when they were available for the asteroids in our sample. This
11
is a departure from the methods in Nugent et al. (2015), which employed MPC database H
and G values as no large replacement dataset was available at that time. Unless specified
otherwise, G is assumed to be 0.15 ± 0.1 mag, and the error in H is assumed to be ±0.3
mag.
3.3. Diameter and albedo calculations
The effective diameter d of each asteroid and geometric optical albedo pv were then
calculated from the resulting verified, high-quality minor planet measurements using the
Near-Earth Asteroid Thermal Model (NEATM; Harris 1998). The implementation used
in this work is detailed in Mainzer et al. (2011c). It assumes a spherical object with no
rotation, no nightside emission, and a temperature distribution given by:
T (θ) = Tmax cos1/4(θ)
for 0 ≤ θ ≤ π/2
(1)
where θ is the angular distance from the sub-solar point. Tmax is the sub-solar temperature,
defined as:
Tmax =
ησ
(cid:18)(1 − A)S
(cid:19)1/4
(2)
where A is the bolometric Bond albedo, S is the solar flux at the asteroid, η is the
beaming parameter, is the emissivity, and σ is the Stefan-Boltzmann constant. The
beaming parameter η adjusts the temperature distribution, and variation of η can be due to
non-spherical shapes, rotation rates, spin pole orientation with respect to observer, surface
thermal inertia, phase effects, etc.
After a best-fit diameter is found, twenty-five Monte Carlo trials were run to evaluate
the errors introduced by the uncertainty in the flux measurements. The corresponding
uncertainties in diameter and albedo, along with the H and G values used as inputs to the
thermal model, are reported in Tables 4 and 5.
12
The NEOWISE survey cadence observes each object over ∼ 1.5 days on average,
and sometimes re-observes an object ∼ 3 to ∼ 6 months later at a different distance and
viewing geometry. These separate epochs, defined as observations separated by > 10 days,
the typical amount of time for viewing geometry of NEOs to change significantly, were fit
separately.
NEAs were treated differently than Mars-crossing and main belt asteroids, because of
the different characteristics of the populations and different phase angles as demonstrated
in Mainzer et al. (2011b) and Masiero et al. (2011). In most cases, NEAs were fit with
η = 1.4 ± 0.5. If both bands were thermally dominated, a beaming parameter was fit. A
ratio of pIR/pV = 1.6 ± 1.0 was assumed for NEAs. Mars-crossing and main belt asteroids
were fit with η = 0.95 ± 0.2, and the ratio of pIR/pV was taken to be 1.5 ± 0.1 in most
cases. These assumptions were necessary because if only one thermally dominated band
is available, a beaming parameter cannot be fit; similarly, with only the 3.4 and 4.6 µm
bands, we cannot fit pIR because there is not enough information to constrain it.
As noted in Tables 4 and 5, some objects were fit with alternative beaming
parameters and pIR/pV ratios. In rare cases the standard assumption of η = 1.4 ± 0.5
and η = 0.95 ± 0.2 for NEAs and MBAs, respectively, lead to poor fits. Poor fits are
indicated by abs(Hobserved − Hmodeled) > 0.5 or unphysical values of pV , generally taken to
be pV <∼ 0.02, pV >∼ 0.6 ). In cases where a poor fit is obtained, we use the constraints
on H magnitude errors and physical limits on albedo to exclude unphysical results, and
rule out certain beaming and pIR/pV values. A series of broadly-spaced beaming values (in
increments of 0.2) and pIR/pV ratios (in increments of 0.5) were tried; in these few cases,
the associated errors were increased. These spacings were chosen so that in most cases only
a single pair of beaming and pIR/pV ratios would produce a good fit.
NEATM is only effective if at least one of the wavelength bands employed is dominated
by thermal emission. Therefore, any object found to have < 75% thermally emitted light
13
(generally the cooler outer main belt objects) in both bands was removed from the results.
This determination is made after an initial fit to the object is completed and estimates of
thermally emitted and reflected light can be computed.
4. Results
Thermal fit results for NEAs are presented in Table 4; Table 5 contains the fit results
for Mars-crossing and main belt asteroids. When objects were observed at multiple epochs,
a measurement of diameter and albedo is given for each epochs.
Some asteroids have diameters and albedos calculated from earlier NEOWISE
measurements (Mainzer et al. 2011b; Masiero et al. 2011; Mainzer et al. 2012a; Masiero
et al. 2012; Nugent et al. 2015). Figure 3 is a histogram of the diameters and albedos for
these objects measured from Reactivation Year 2 data and previous work. This figure also
compares the results of the corrected H and G values from Williams (2012). Although
distributions of diameter and albedo for this work are comparable to previous NEOWISE
results, the incorporation of the revised H and G values does shift the albedo distribution
towards slightly lower values. The implementation of the Williams (2012) H and G values
did not change the diameters of the ensemble of NEAs or other asteroids in a statistically
significant way (see Figure 3).
When possible, diameters calculated from this work were compared to diameters
calculated by independent methods (Figure 4). Twenty-three objects have diameters
calculated via stellar occultations (Shevchenko & Tedesco 2006), eleven have radar-derived
shapes (Benner et al. 2015), and two, (951) Gaspra and (253) Mathilde, were observed
by spacecraft and had shape and size determined from resulting images (Thomas et al.
1994, 1999). These comparison cases were not preselected on light curve amplitude.
When three-dimensional shapes were known, comparison was made to the average of
14
the length of each axis. As illustrated in Figure 4, a Gaussian fit to a histogram of
(DN EOW ISE − Dref erence)/DN EOW ISE gives σ = 20%, and a Gaussian fit to a histogram of
(pv−N EOW ISE − pv−ref erence)/pv−N EOW ISE gives σ = 40%.
We report the Gaussian-fit 1 − σ uncertainty of 20% on diameter, and 40% on albedo,
based on the comparison to diameter measurements made with other techniques known to
produce highly accurate diameters. This encompasses the systematic uncertainties in the
comparison measurements (radar, stellar occultation, and spacecraft measurements), the
range of ways that actual objects do not precisely match with the assumptions of NEATM,
as well as the color corrections derived for the WISE filters (Wright et al. 2010).
The diameters and albedos of NEOWISE Reactivation discoveries are compared with
the diameters and albedos of objects detected during Reactivation operations in Figure 5.
NEOWISE continues to discover large objects (> 100 m), as well as low-albedo objects.
4.1. Potentially Hazardous Asteroids
Potentially Hazardous Asteroids (PHAs) have been defined as objects with H ≤ 22.0
mag and a Minimum Orbit Intersection Distance (MOID) of 0.05 AU. The MOID is a
measurement of the smallest distance between two orbits (Sitarski 1968; Gronchi 2005).
Since many NEAs do not have measured diameters, the H limit was used as a proxy for
size. An object with pV =∼ 0.14 and H = 22.0 mag corresponds to an object ∼ 140 m in
diameter.
Using the PHA definition as defined by H limit, five NEOWISE Reactivation Year 2
discoveries are considered PHAs. However, eight NEOWISE Reactivation Year 2 discoveries
are larger than 140 m in diameter and have a MOID ≤ 0.05 AU, and therefore should be
classified as PHAs. With the availability of more diameter measurements of NEAs from
NEOWISE, the Spitzer Space Telescope (Trilling et al. 2010), and ground-based facilities,
15
sizes should be taken into consideration when designating PHAs as suggested in Mainzer
et al. (2012b). The fraction of PHAs within the NEOWISE NEA discoveries remains
virtually constant across Year 1 and Year 2 of the Reactivation mission, and is nearly a
factor of three higher than ground-based surveys3.
4.2. NHATS
The Near-Earth Object Human Space Flight Accessible Targets Study (NHATS,
Barbee et al. 2013) aims to identify the asteroids that would be most accessible to a
crewed mission to an asteroid4. NHATS-compliant targets must pass a series of restrictions,
including Earth departure dates before 31 Dec 2040, total mission ∆V ≤ 12 km s−1, and a
minimum NEA stay time of 8 days. Many of these objects do not have measured physical
properties. The NEOWISE Year 2 Reactivation mission measured diameters and albedos
for eight objects (Table 6). Two, (35107) 1991 VH and (363505) 2003 UC20, were observed
during Reactivation Year 1.
5. Conclusion
NEOWISE continues its mission to discover, track, and characterize minor planets.
This release of diameters and albedos for 9,092 asteroids measured using NEOWISE Year 2
observations increases the total number of asteroids with measured diameters and albedos
by 1,440, enabling further studies of NEAs and other asteroids by the scientific community
and provides multi-epoch infrared observations that support more detailed thermophysical
modeling studies. Comparison to diameters measured by other methods shows that
3http://neo.jpl.nasa.gov/stats/
4http://neo.jpl.nasa.gov/nhats/
16
measured diameters continue to be accurate to ∼ 20 + % during the Year 2 Reactivation
mission. NEOWISE continues to preferentially discover large (> 100m), low-albedo NEOs.
6. Acknowledgments
The authors thank the reviewer, Valerio Carruba, for his careful review that improved
the quality of this manuscript.
This publication makes use of data products from the Wide-field Infrared Survey
Explorer, which is a joint project of the University of California, Los Angeles, and
JPL/California Institute of Technology, funded by NASA. This publication also makes use
of data products from NEOWISE, which is a project of the JPL/California Institute of
Technology, funded by the Planetary Science Division of NASA. The JPL High-Performance
Computing Facility used for our simulations is supported by the JPL Office of the CIO.
This research has made use of the NASA/ IPAC Infrared Science Archive, which
is operated by the Jet Propulsion Laboratory, California Institute of Technology, under
contract with the National Aeronautics and Space Administration.
This project used data obtained with the Dark Energy Camera (DECam), which was
constructed by the Dark Energy Survey (DES) collaboration. Funding for the DES Projects
has been provided by the U.S. Department of Energy, the U.S. National Science Foundation,
the Ministry of Science and Education of Spain, the Science and Technology Facilities
Council of the United Kingdom, the Higher Education Funding Council for England, the
National Center for Supercomputing Applications at the University of Illinois at Urbana-
Champaign, the Kavli Institute of Cosmological Physics at the University of Chicago, the
Center for Cosmology and Astro-Particle Physics at the Ohio State University, the Mitchell
Institute for Fundamental Physics and Astronomy at Texas A&M University, Financiadora
de Estudos e Projetos, Funda¸cao Carlos Chagas Filho de Amparo `a Pesquisa do Estado do
17
Rio de Janeiro, Conselho Nacional de Desenvolvimento Cient´ıfico e Tecnol´ogico and the
Minist´erio da Ciencia, Tecnologia e Inovacao, the Deutsche Forschungsgemeinschaft, and
the Collaborating Institutions in the Dark Energy Survey. The Collaborating Institutions
are Argonne National Laboratory, the University of California at Santa Cruz, the University
of Cambridge, Centro de Investigaciones En´ergeticas, Medioambientales y Tecnol´ogicas-
Madrid, the University of Chicago, University College London, the DES-Brazil Consortium,
the University of Edinburgh, the Eidgenossische Technische Hochschule (ETH) Zurich,
Fermi National Accelerator Laboratory, the University of Illinois at Urbana-Champaign,
the Institut de Ci`encies de l'Espai (IEEC/CSIC), the Institut de F´ısica d'Altes Energies,
Lawrence Berkeley National Laboratory, the Ludwig-Maximilians Universitat Munchen
and the associated Excellence Cluster Universe, the University of Michigan, the National
Optical Astronomy Observatory, the University of Nottingham, the Ohio State University,
the University of Pennsylvania, the University of Portsmouth, SLAC National Accelerator
Laboratory, Stanford University, the University of Sussex, and Texas A&M University.
This work makes use of observations from the LCOGT network.
The authors wish to thank G. Williams, for providing the corrected H and G values
from his dissertation used in this work.
This publication makes use of observations obtained at the Gemini Observatory,
which is operated by the Association of Universities for Research in Astronomy, Inc.,
under a cooperative agreement with the NSF on behalf of the Gemini partnership: the
National Science Foundation (United States), the National Research Council (Canada),
CONICYT (Chile), Ministerio de Ciencia, Tecnolog´ıa e Innovaci´on Productiva (Argentina),
and Minist´erio da Ciencia, Tecnologia e Inova¸cao (Brazil). Observing Program IDs:
GS-2015A-LP-3, GS-2015B-LP-3.
18
Fig. 3.- Comparison between asteroid diameters (top) and albedos (bottom) measured
in this work with H and G values from the MPC (blue), diameters for the same objects
measured in this work with revised H and G values from Williams (2012) (black), and
diameters for the same objects measured using previous NEOWISE measurements, which
employed H and G values from the MPC (green). The bimodal structure of the albedo
distribution is due to the populations of bright S-type (pV = 0.25) and dark C-type (pV =
0.06) objects in the main belt.
19
Fig. 4.- Top: Comparison of diameters and albedos derived via radar, stellar occultations,
and spacecraft flybys to the values calculated in this paper. The dashed red line shows a 1:1
relation. Bottom: Histograms of the fractional differences between the NEOWISE diameters
(%∆d, left) and albedos (%∆pV , right) and those derived from other methods. Dashed red
line is best-fit Gaussian, with the fitted σ given in the legends.
20
Fig. 5.- Diameters and albedos from NEOWISE measurements of previously known NEAs
(teal circles) and NEOWISE NEA discoveries (black squares) made during years 1 and 2 of
the Reactivation. NEOWISE continues to detect large objects > 100 m, and many discoveries
are dark. Error bars on previously known objects were omitted for clarity.
21
REFERENCES
Barbee, B. W., Abell, P. A., Adamo, D. R., et al. 2013, 2013 IAA Planetary Defense
Conference Proceedings
Bauer, J. M., Grav, T., Blauvelt, E., et al. 2013, The Astrophysical Journal, 773, 22
Bauer, J. M., Kramer, E., Mainzer, A. K., et al. 2012a, The Astrophysical Journal, 758, 18
Bauer, J. M., Mainzer, A. K., Grav, T., et al. 2012b, The Astrophysical Journal, 747, 49
Bauer, J. M., Stevenson, R., Kramer, E., et al. 2015, The Astrophysical Journal, 814, 85
Bauer, J. M., Walker, R. G., Mainzer, A. K., et al. 2011, The Astrophysical Journal, 738,
171
Benner, L., Busch, M. B., Giorgini, J. D., Taylor, P. A., & Margot, J. L. 2015, Asteroids IV
Bowell, E., Hapke, B., Domingue, D., et al. 1989, in Asteroids II, 524–556
Brown, T. M., Baliber, N., Bianco, F. B., et al. 2013, Publications of the Astronomical
Society of the Pacific, 125, 1031
Bus, S. J. & Binzel, R. P. 2002, Icarus, 158, 146
Carruba, V., Domingos, R. C., Nesvorn´y, D., et al. 2013, Monthly Notices of the Royal
Astronomical Society, 433, 2075
Christensen, E. J., Carson Fuls, D., Gibbs, A. R., et al. 2015, in AAS/Division for Planetary
Sciences Meeting Abstracts, Vol. 47, AAS/Division for Planetary Sciences Meeting
Abstracts, 308.19
Cutri, R. M., Mainzer, A., Conrow, T., et al. 2015, 1, explana-
tory Supplement to the NEOWISE Data Release Products,
http://wise2.ipac.caltech.edu/docs/release/neowise/expsup
22
Cutri, R. M., Wright, E. L., Conrow, T., et al. 2012, 1,
http://wise2.ipac.caltech.edu/docs/release/allsky/expsup/sec8 1.html
Flaugher, B., Diehl, H. T., Honscheid, K., et al. 2015, AJ, 150, 150
Gradie, J. & Tedesco, E. 1982, Science, 216, 1405
Grav, T., Mainzer, A. K., Bauer, J., et al. 2012a, The Astrophysical Journal, 744, 197
-. 2011, The Astrophysical Journal, 742, 40
Grav, T., Mainzer, A. K., Bauer, J. M., Masiero, J. R., & Nugent, C. R. 2012b, The
Astrophysical Journal, 759, 49
Grav, T. et al. 2015, The Astrophysical Journal, 809, 3
Gronchi, G. F. 2005, Celestial Mechanics and Dynamical Astronomy, 93, 295
Harris, A. W. 1998, Icarus, 131, 291
Hook, I. M., Jørgensen, I., Allington-Smith, J. R., et al. 2004, The Publications of the
Astronomical Society of the Pacific, 116, 425
Jedicke, R., Larsen, J., & Spahr, T. 2002, Observational Selection Effects in Asteroid
Surveys, ed. W. F. Bottke, Jr., A. Cellino, P. Paolicchi, & R. P. Binzel, 71–87
Kramer, E. A., Bauer, J. M., Fern´andez, Y. R., et al. 2015, in AAS/Division for Planetary
Sciences Meeting Abstracts, Vol. 47, AAS/Division for Planetary Sciences Meeting
Abstracts, 506.09
Mainzer, A., Bauer, J., Cutri, R. M., et al. 2014a, The Astrophysical Journal, 792, 30
Mainzer, A., Bauer, J., Grav, T., et al. 2011a, The Astrophysical Journal, 731, 53
-. 2014b, The Astrophysical Journal, 784, 110
23
Mainzer, A., Grav, T., Bauer, J., et al. 2011b, The Astrophysical Journal, 743, 156
Mainzer, A., Grav, T., Masiero, J., et al. 2012a, The Astrophysical Journal Letters, 760,
L12
-. 2012b, The Astrophysical Journal, 752, 110
-. 2011c, The Astronomical Journal, 736, 100
-. 2011d, The Astrophysical Journal Letters, 737, L9
Masiero, J., DeMeo, F., Kasuga, T., & Parker, A. H. 2015a, ArXiv e-prints
Masiero, J. R., Grav, T., Mainzer, A. K., et al. 2014, The Astrophysical Journal, 791, 121
Masiero, J. R., Mainzer, A. K., Bauer, J. M., et al. 2013, The Astrophysical Journal, 770, 7
Masiero, J. R., Mainzer, A. K., Grav, T., et al. 2011, The Astrophysical Journal, 741, 68
-. 2012, The Astrophysical Journal Letters, 759, L8
Masiero, J. R. et al. 2015b, The Astrophysical Journal, 809, 179
McMillan, R. S. 2007, in IAU Symposium, Vol. 236, IAU Symposium, ed. G. B. Valsecchi,
D. Vokrouhlick´y, & A. Milani, 329–340
Milani, A., Cellino, A., Knezevi´c, Z., et al. 2014, Icarus, 239, 46
Nugent, C. R., Mainzer, A., Masiero, J., et al. 2015, The Astrophysical Journal, 814, 117
Pravec, P., Harris, A. W., Kusnir´ak, P., Gal´ad, A., & Hornoch, K. 2012, Icarus, 221, 365
Shevchenko, V. G. & Tedesco, E. F. 2006, Icarus, 184, 211
Sitarski, G. 1968, Acta Astronomica, 18, 171
24
Tedesco, E. F., Noah, P. V., Noah, M., & Price, S. D. 2002, The Astronomical Journal, 123,
1056
Tholen, D. J., Mainzer, A. K., Bauer, J. M., et al. 2014, Minor Planet Electronic Circulars,
145
Thomas, P. C., Veverka, J., Bell, J. F., et al. 1999, Icarus, 140, 17
Thomas, P. C., Veverka, J., Simonelli, D., et al. 1994, Icarus, 107, 23
Trilling, D. E., Mueller, M., Hora, J. L., et al. 2010, The Astronomical Journal, 140, 770
Veres, P., Jedicke, R., Fitzsimmons, A., et al. 2015, Icarus, 261, 34
Walsh, K. J., Delb´o, M., Bottke, W. F., Vokrouhlick´y, D., & Lauretta, D. S. 2013, Icarus,
225, 283
Warner, B. D., Harris, A. W., & Pravec, P. 2009, Icarus, 202, 134
Williams, G. V. 2012, PhD thesis, Open University UK
Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, The Astronomical Journal,
140, 1868
Zellner, B. 1979, Asteroid taxonomy and the distribution of the compositional types, ed.
T. Gehrels (The University of Arizona Press), 783–806
This manuscript was prepared with the AAS LATEX macros v5.2.
25
Table 3. NEOWISE magnitudes for the NEAs modeled in this paper. Listed are
MPC-packed name, the time of the observation in modified Julian date (MJD), and the
magnitude in the 3.4µm (W1) and 4.6µm bands (W2). Non-detections at a particular
wavelength represent 95% confidence limits (Cutri et al. 2012). Observations for the first
two objects only are shown; the remainder are available in electronic format through the
journal website.
Object
MJD
W1 (mag)
W2 (mag)
01580
57059.5362951
01580
57059.7989937
01580
57059.9304065
01580
57060.0618195
01580
57162.4156505
01580
57162.4812294
01580
57162.5468084
01580
57162.6780938
01580
57162.8092518
01580
57163.2684322
01580
57163.5308722
01580
57163.5309995
01580
57163.7933154
01580
57163.9246007
01580
57164.1213378
01580
57164.3837815
01580
57164.7772554
01580
57164.8429617
01580
57164.9741197
01580
57165.1052777
01580
57165.2364356
01580
57165.2365629
01580
57189.9632607
01580
57190.0288402
01580
57190.0944191
01580
57190.1601254
01580
57190.2257043
01580
57190.2912832
01580
57190.3568621
01620
57259.6667067
01620
57259.7979924
01620
57259.9291508
14.638 ± 0.074
14.540 ± 0.068
14.564 ± 0.076
14.525 ± 0.084
11.563 ± 0.017
11.507 ± 0.017
11.586 ± 0.018
11.527 ± 0.019
11.449 ± 0.017
11.724 ± 0.018
11.825 ± 0.021
11.836 ± 0.021
11.879 ± 0.019
11.808 ± 0.021
11.317 ± 0.017
11.406 ± 0.021
11.442 ± 0.021
11.566 ± 0.021
11.507 ± 0.019
11.417 ± 0.020
11.388 ± 0.026
11.432 ± 0.021
12.328 ± 0.023
12.292 ± 0.023
12.292 ± 0.026
12.022 ± 0.028
12.318 ± 0.028
12.307 ± 0.022
12.424 ± 0.025
14.743 ± 0.082
14.789 ± 0.104
26
15.247 ± 0.130
11.338 ± 0.024
11.339 ± 0.024
11.274 ± 0.025
11.290 ± 0.028
8.478 ± 0.015
8.436 ± 0.012
8.497 ± 0.013
8.436 ± 0.014
8.388 ± 0.013
8.624 ± 0.015
8.695 ± 0.014
8.443 ± 0.014
8.803 ± 0.013
8.737 ± 0.013
8.222 ± 0.013
8.286 ± 0.015
8.373 ± 0.015
8.512 ± 0.013
8.381 ± 0.012
8.371 ± 0.014
8.330 ± 0.014
8.316 ± 0.012
9.224 ± 0.014
9.209 ± 0.013
9.204 ± 0.015
8.925 ± 0.012
9.243 ± 0.014
9.198 ± 0.014
9.268 ± 0.015
12.908 ± 0.061
12.701 ± 0.055
13.368 ± 0.099
Table 3-Continued
Object
MJD
W1 (mag)
W2 (mag)
01620
57260.0603091
01620
57260.2570466
01620
57260.2571739
01620
57260.3227531
01620
57260.3883323
01620
57260.4539115
01620
57260.5850698
01620
57260.7818073
01620
57260.913093
01620
57261.1754097
15.713 ± 0.177
15.336 ± 0.134
15.263 ± 0.125
14.641 ± 0.075
15.517 ± 0.150
14.675 ± 0.079
14.916 ± 0.091
14.934 ± 0.108
15.390 ± 0.135
15.203 ± 0.115
13.865 ± 0.138
13.066 ± 0.071
13.355 ± 0.092
12.867 ± 0.067
13.546 ± 0.158
12.747 ± 0.067
13.053 ± 0.069
12.796 ± 0.060
13.504 ± 0.114
13.108 ± 0.070
27
Table 4. Measured diameters (d) and albedos (pV ) of near-Earth asteroids observed
during the NEOWISE Year 2 mission. Asteroids may be identified by numbers, provisional
designations, or via the MPC packed format. Magnitude H, slope parameter G, and
beaming η used are given. The numbers of observations used in the 3.4 µm (nW 1) and 4.6
µm (nW 2) wavelengths are also reported, along with the amplitude of the 4.6 µm light
curve (W2 amp., in mag).
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
1580
1580
1580
1620
1685
1980
1980
1980
2062
2063
3691
4055
4183
5646
5646
5646
5731
5828
7335
7350
7889
9202
9400
11066
11405
21088
22099
23606
26817
35107
38086
38091
01580
14.90
0.12
01580
14.90
0.12
01580
14.90
0.12
01620
15.41
0.24
01685
14.45
0.24
01980
13.87
0.24
01980
13.87
0.24
01980
13.87
0.24
02062
17.30
0.24
02063
17.37
0.24
03691
14.98
0.24
04055
14.99
0.43
04183
14.35
0.24
05646
15.45
0.24
05646
15.45
0.24
05646
15.45
0.24
05731
15.53
0.12
05828
16.30
0.24
07335
17.82
0.24
07350
17.21
0.24
07889
15.31
0.24
09202
16.16
0.24
09400
14.98
0.24
11066
15.36
0.24
11405
15.37
0.24
21088
14.47
0.24
22099
18.19
0.24
23606
18.37
0.24
26817
19.11
0.24
35107
17.02
0.24
38086
17.63
0.24
38091
16.61
0.24
7.91 ± 0.08
4.19 ± 0.06
5.37 ± 0.04
1.96 ± 0.06
3.91 ± 0.08
4.36 ± 0.10
4.31 ± 0.18
4.47 ± 0.14
0.73 ± 0.03
1.03 ± 0.03
2.08 ± 0.09
3.21 ± 0.16
3.73 ± 0.15
2.45 ± 0.06
2.50 ± 0.05
2.51 ± 0.05
6.51 ± 3.14
1.43 ± 0.06
0.73 ± 0.02
1.92 ± 0.03
1.82 ± 0.08
1.51 ± 0.05
3.69 ± 0.05
2.10 ± 0.09
3.62 ± 0.05
2.79 ± 0.10
0.52 ± 0.11
0.87 ± 0.01
1.17 ± 0.53
0.91 ± 0.03
0.64 ± 0.19
2.49 ± 0.03
0.03 ± 0.01
0.11 ± 0.02
0.07 ± 0.01
0.32 ± 0.04
0.19 ± 0.02
0.26 ± 0.03
0.27 ± 0.05
0.25 ± 0.06
0.39 ± 0.05
0.19 ± 0.03
0.42 ± 0.11
0.17 ± 0.03
0.23 ± 0.04
0.19 ± 0.03
0.19 ± 0.03
0.18 ± 0.03
0.03 ± 0.03
0.26 ± 0.05
0.25 ± 0.04
0.06 ± 0.01
0.40 ± 0.04
0.27 ± 0.04
0.13 ± 0.02
0.29 ± 0.04
0.10 ± 0.02
0.37 ± 0.06
0.35 ± 0.22
0.11 ± 0.02
0.03 ± 0.08
0.33 ± 0.04
0.39 ± 0.25
28
0.06 ± 0.01
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.00 ± 0.51
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.40
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.50
1.40 ± 0.00
2.31 ± 0.10
1.03 ± 0.10
1.61 ± 0.10
1.78 ± 0.10
1.68 ± 0.10
1.83 ± 0.10
1.94 ± 0.10
1.97 ± 0.10
1.46 ± 0.10
0.93 ± 0.10
1.19 ± 0.10
1.70 ± 0.10
1.21 ± 0.10
1.32 ± 0.10
1.28 ± 0.10
1.41 ± 0.10
1.60 ± 0.10
1.29 ± 0.10
1.52 ± 0.10
1.06 ± 0.10
1.27 ± 0.10
1.29 ± 0.10
2.13 ± 0.10
1.83 ± 0.10
2.67 ± 0.10
1.63 ± 0.10
1.60 ± 0.10
2.46 ± 0.10
1.60 ± 0.10
1.41 ± 0.10
1.60 ± 0.10
2.30 ± 0.10
0.07
0.34
0.58
1.16
1.04
0.99
1.36
1.46
0.25
0.24
0.73
0.91
0.33
0.36
0.66
0.35
0.33
0.38
0.78
0.12
0.77
0.18
0.20
0.49
0.25
0.45
0.81
1.67
0.41
0.93
0.71
0.97
4
7
18
13
10
18
31
27
6
5
21
6
9
39
74
46
0
6
8
4
13
5
10
5
13
12
0
60
0
27
0
12
4
7
18
13
11
18
31
29
6
5
21
7
9
39
78
47
8
6
8
5
13
6
10
5
13
12
4
60
6
27
6
13
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
52750
52750
53430
68216
68278
85275
85713
85804
85804
85804
85989
86067
86667
86829
88263
88710
90367
90403
90416
52750
16.74
0.24
52750
16.74
0.24
53430
16.85
0.24
68216
16.66
0.24
68278
18.50
0.12
85275
16.45
0.24
85713
16.01
0.24
85804
15.47
0.24
85804
15.47
0.24
85804
15.47
0.24
85989
17.03
0.24
86067
16.52
0.24
86667
17.59
0.24
86829
16.11
0.24
88263
15.73
0.24
88710
18.14
0.34
90367
18.11
0.24
90403
17.78
0.24
90416
18.58
0.24
100756
A0756
16.48
0.24
105140
A5140
15.81
0.24
108519
A8519
17.94
0.24
112985
B2985
15.66
0.24
112985
B2985
15.66
0.24
137084
D7084
16.52
0.24
137805
D7805
16.77
0.24
137925
D7925
16.25
0.24
140288
E0288
16.84
0.24
140288
E0288
16.84
0.24
141484
E1484
16.64
0.24
141484
E1484
16.64
0.24
142040
E2040
16.31
0.24
0.96 ± 0.04
0.93 ± 0.32
1.34 ± 0.45
0.99 ± 0.04
1.46 ± 0.53
2.50 ± 0.99
3.03 ± 1.49
2.27 ± 0.06
2.19 ± 0.04
2.84 ± 0.04
1.60 ± 0.59
1.50 ± 0.49
0.74 ± 0.02
1.81 ± 0.05
5.10 ± 1.86
0.75 ± 0.29
2.06 ± 1.14
0.57 ± 0.17
0.98 ± 0.02
1.81 ± 0.04
1.97 ± 0.05
1.43 ± 0.54
3.65 ± 0.04
5.12 ± 0.03
1.23 ± 0.04
2.24 ± 0.03
1.36 ± 0.04
1.26 ± 0.03
1.21 ± 0.45
1.00 ± 0.04
1.02 ± 0.03
1.26 ± 0.04
1.40 ± 0.00
1.40 ± 0.51
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.40
1.40 ± 0.42
1.40 ± 0.50
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.42
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.38
1.40 ± 0.52
1.40 ± 0.59
1.40 ± 0.49
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.50
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.55 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.30 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.24 ± 0.10
1.29 ± 0.10
1.94 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.38 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.34 ± 0.10
1.80 ± 0.10
1.30 ± 0.10
1.60 ± 0.10
2.46 ± 0.10
2.80 ± 0.10
1.10 ± 0.10
1.40 ± 0.10
1.27 ± 0.10
1.03 ± 0.10
1.60 ± 0.10
1.44 ± 0.10
1.59 ± 0.10
1.71 ± 0.10
0.39 ± 0.06
0.41 ± 0.25
0.18 ± 0.17
0.39 ± 0.06
0.03 ± 0.03
0.07 ± 0.06
0.08 ± 0.12
0.22 ± 0.04
0.24 ± 0.02
0.14 ± 0.02
0.11 ± 0.17
0.19 ± 0.14
0.29 ± 0.05
0.19 ± 0.03
0.03 ± 0.04
0.18 ± 0.14
0.02 ± 0.07
0.42 ± 0.24
0.07 ± 0.01
0.14 ± 0.02
0.22 ± 0.04
0.06 ± 0.09
0.07 ± 0.01
0.04 ± 0.01
0.29 ± 0.04
0.07 ± 0.01
0.30 ± 0.05
0.20 ± 0.04
0.22 ± 0.23
0.39 ± 0.05
0.37 ± 0.04
0.34 ± 0.04
29
0.73
0.19
0.78
0.38
0.52
0.53
0.37
0.17
0.21
0.58
0.39
0.54
0.76
0.13
0.93
0.99
0.48
0.49
0.07
1.16
0.58
0.52
0.12
0.15
0.48
0.31
0.29
0.34
0.74
0.26
0.39
0.36
22
0
0
8
0
0
0
11
13
21
0
0
4
6
0
0
0
0
5
15
8
0
11
18
4
17
4
8
0
8
16
43
24
5
10
8
23
14
6
11
14
22
4
11
4
6
18
71
13
5
5
15
9
12
11
18
4
17
5
8
5
9
16
44
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
152679
F2679
16.43
0.12
152978
F2978
19.73
0.24
152978
F2978
19.73
0.24
153195
F3195
17.94
0.24
153195
F3195
17.94
0.24
154807
F4807
18.72
0.24
155110
F5110
17.66
0.24
155110
F5110
17.66
0.24
159459
F9459
15.98
0.24
159504
F9504
16.99
0.05
159686
F9686
16.65
0.24
159929
F9929
17.75
0.24
161989
G1989
17.43
0.24
162080
G2080
19.89
0.24
162463
G2463
17.99
0.24
162463
G2463
17.99
0.24
162567
G2567
20.18
0.24
163760
G3760
16.53
0.24
163899
G3899
17.36
0.24
164206
G4206
17.86
0.24
172034
H2034
17.67
0.24
173689
H3689
18.28
0.24
190161
J0161
16.67
0.24
200754
K0754
18.67
0.24
206378
K6378
18.68
0.06
212359
L2359
16.98
0.24
237805
N7805
17.63
0.24
241662
O1662
17.64
0.24
241662
O1662
17.64
0.24
248590
O8590
16.82
0.24
256412
P6412
17.17
0.24
275611
R5611
18.24
0.24
4.18 ± 0.01
0.32 ± 0.07
0.37 ± 0.14
1.32 ± 0.55
1.60 ± 0.59
0.47 ± 0.01
0.68 ± 0.03
0.75 ± 0.32
1.83 ± 0.09
2.31 ± 0.02
1.80 ± 0.03
2.62 ± 1.20
0.64 ± 0.02
0.68 ± 0.32
0.93 ± 0.35
0.98 ± 0.35
0.29 ± 0.12
2.35 ± 0.77
0.80 ± 0.02
1.13 ± 0.55
0.66 ± 0.17
0.73 ± 0.29
3.05 ± 0.02
0.56 ± 0.21
0.37 ± 0.02
1.25 ± 0.40
0.69 ± 0.26
0.91 ± 0.02
0.81 ± 0.26
3.35 ± 1.04
2.90 ± 0.02
1.48 ± 0.01
1.40 ± 0.00
1.40 ± 0.32
1.40 ± 0.47
1.40 ± 0.47
1.40 ± 0.39
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.54
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.00
1.40 ± 0.55
1.40 ± 0.49
1.40 ± 0.43
1.40 ± 0.56
1.40 ± 0.39
1.40 ± 0.00
1.40 ± 0.55
1.40 ± 0.44
1.40 ± 0.55
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.42
1.40 ± 0.57
1.40 ± 0.00
1.40 ± 0.47
0.80 ± 0.40
1.00 ± 0.20
1.40 ± 0.00
7.21 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.69 ± 0.10
1.87 ± 0.10
1.60 ± 0.10
1.11 ± 0.10
1.73 ± 0.10
3.56 ± 0.10
1.60 ± 0.10
1.81 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.35 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.79 ± 0.10
1.60 ± 0.10
1.85 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.56 ± 0.10
1.60 ± 0.10
1.50 ± 0.10
4.06 ± 0.10
2.67 ± 0.10
0.03 ± 0.00
0.23 ± 0.13
0.17 ± 0.14
0.07 ± 0.07
0.05 ± 0.04
0.26 ± 0.03
0.33 ± 0.04
0.27 ± 0.23
0.21 ± 0.03
0.05 ± 0.01
0.12 ± 0.02
0.02 ± 0.08
0.46 ± 0.09
0.04 ± 0.05
0.13 ± 0.21
0.12 ± 0.14
0.18 ± 0.16
0.08 ± 0.08
0.31 ± 0.04
0.10 ± 0.09
0.34 ± 0.25
0.16 ± 0.19
0.04 ± 0.01
0.19 ± 0.15
0.44 ± 0.19
0.18 ± 0.18
0.33 ± 0.22
0.19 ± 0.03
0.24 ± 0.22
0.03 ± 0.03
0.03 ± 0.01
0.04 ± 0.01
30
1.22
0.74
0.57
0.77
0.45
0.92
0.34
0.37
0.38
0.24
0.28
0.34
1.16
0.94
0.85
0.40
0.19
0.40
1.97
0.94
1.17
0.17
0.36
0.29
0.75
0.33
1.47
0.38
0.44
0.27
0.57
0.70
29
0
0
0
0
14
6
0
5
9
8
0
64
0
0
0
0
0
24
0
0
0
28
0
20
0
0
10
0
0
15
24
29
8
5
14
7
15
6
5
6
9
9
8
69
20
21
6
5
9
24
8
24
5
30
5
20
9
45
10
8
13
15
25
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
276049
R6049
16.50
0.12
276049
R6049
16.50
0.12
276786
R6786
18.11
0.24
285331
S5331
18.47
0.24
285331
S5331
18.47
0.24
294739
T4739
17.39
0.24
297274
T7274
16.90
0.24
303450
U3450
20.86
0.24
307493
U7493
18.95
0.24
311554
V1554
18.80
0.24
326388 W6388
18.26
0.24
337248
X7248
20.00
0.15
337248
X7248
20.00
0.15
345646
Y5646
19.90
0.15
355770
Z5770
18.40
0.15
363027
a3027
19.50
0.15
363027
a3027
19.50
0.15
363505
a3505
18.10
0.15
373135
b3135
19.50
0.15
381906
c1906
17.90
0.15
385186
c5186
17.70
0.15
385186
c5186
17.70
0.15
401857
e1857
16.10
0.15
401857
e1857
16.10
0.15
401925
e1925
18.40
0.15
413123
f3123
19.00
0.15
413123
f3123
19.00
0.15
413192
f3192
16.80
0.15
413192
f3192
16.80
0.15
414287
f4287
17.70
0.15
414772
f4772
19.00
0.15
415711
f5711
19.00
0.15
2.24 ± 0.02
4.71 ± 2.84
1.72 ± 0.68
0.66 ± 0.01
0.65 ± 0.01
0.74 ± 0.21
1.21 ± 0.03
0.18 ± 0.06
1.43 ± 0.09
0.38 ± 0.02
1.15 ± 0.01
0.85 ± 0.06
0.61 ± 0.26
0.41 ± 0.01
1.20 ± 0.49
0.58 ± 0.27
0.69 ± 0.25
1.88 ± 0.01
1.05 ± 0.36
0.52 ± 0.10
0.81 ± 0.02
0.97 ± 0.29
4.28 ± 0.04
3.90 ± 1.87
0.48 ± 0.10
1.22 ± 0.50
1.26 ± 0.52
2.78 ± 0.02
2.16 ± 0.02
1.97 ± 0.74
1.00 ± 0.68
0.35 ± 0.10
1.00 ± 0.20
1.00 ± 0.60
1.40 ± 0.45
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.43
1.22 ± 0.06
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.09
1.40 ± 0.49
1.40 ± 0.00
1.40 ± 0.48
1.40 ± 0.54
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.38
1.40 ± 0.39
1.40 ± 0.00
1.40 ± 0.41
1.40 ± 0.00
1.40 ± 0.47
1.40 ± 0.37
1.40 ± 0.45
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.41
1.00 ± 0.73
1.40 ± 0.46
0.23 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.04 ± 0.10
1.96 ± 0.10
1.60 ± 0.10
1.01 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.07 ± 0.10
0.70 ± 0.10
4.79 ± 0.10
1.60 ± 0.10
8.52 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.63 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.04 ± 0.10
1.60 ± 0.10
2.28 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.31 ± 0.10
1.49 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.09 ± 0.02
0.02 ± 0.04
0.03 ± 0.07
0.17 ± 0.02
0.17 ± 0.02
0.36 ± 0.19
0.21 ± 0.03
0.24 ± 0.12
0.02 ± 0.00
0.36 ± 0.05
0.07 ± 0.01
0.02 ± 0.01
0.05 ± 0.06
0.12 ± 0.02
0.05 ± 0.08
0.08 ± 0.11
0.06 ± 0.07
0.03 ± 0.00
0.03 ± 0.01
0.45 ± 0.27
0.22 ± 0.03
0.16 ± 0.15
0.03 ± 0.01
0.04 ± 0.07
0.34 ± 0.19
0.03 ± 0.05
0.03 ± 0.08
0.04 ± 0.01
0.07 ± 0.01
0.04 ± 0.04
0.04 ± 0.06
0.37 ± 0.26
31
0.24
0.37
0.85
0.32
0.65
0.35
0.64
0.62
0.34
0.72
0.39
1.01
0.34
0.42
1.68
0.30
0.19
0.70
0.88
0.43
0.57
0.43
0.43
0.78
0.43
0.89
0.44
0.33
0.84
0.57
0.95
0.66
11
0
0
6
10
0
5
0
10
7
11
6
0
6
0
0
0
26
0
0
9
0
6
0
0
0
0
23
114
0
0
0
11
12
12
6
10
7
5
7
10
8
12
6
7
7
57
4
6
28
4
5
9
20
6
11
5
29
12
25
118
9
5
9
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
415986
f5986
18.10
0.15
415986
f5986
18.10
0.15
416071
f6071
17.90
0.15
417264
f7264
17.20
0.15
417264
f7264
17.20
0.15
417264
f7264
17.20
0.15
418797
f8797
19.50
0.15
422699
g2699
18.30
0.15
424089
g4089
17.70
0.15
424392
g4392
21.90
0.15
428223
g8223
16.10
0.15
429746
g9746
17.30
0.15
431107
h1107
17.70
0.15
433953
h3953
20.90
0.15
433992
h3992
18.00
0.15
434096
h4096
18.00
0.15
434633
h4633
20.90
0.15
434633
h4633
20.90
0.15
436671
h6671
18.00
0.15
437879
h7879
17.70
0.15
437994
h7994
17.30
0.15
438990
h8990
18.30
0.15
439889
h9889
20.10
0.15
442605
i2605
19.10
0.15
442742
i2742
17.60
0.15
443806
i3806
22.00
0.15
443880
i3880
19.40
0.15
443923
i3923
17.40
0.15
445025
i5025
17.50
0.15
445305
i5305
19.90
0.15
450159
j0159
18.90
0.15
453687
j3687
19.30
0.15
1.07 ± 0.54
1.08 ± 0.29
0.80 ± 0.01
1.93 ± 0.02
1.93 ± 0.01
2.72 ± 1.26
0.79 ± 0.23
0.62 ± 0.24
2.31 ± 0.72
0.24 ± 0.10
2.53 ± 0.82
1.27 ± 0.50
1.27 ± 0.44
0.26 ± 0.09
0.88 ± 0.02
0.53 ± 0.02
0.31 ± 0.12
0.44 ± 0.14
2.16 ± 0.02
2.24 ± 0.95
0.80 ± 0.21
0.82 ± 0.37
0.59 ± 0.18
0.44 ± 0.11
2.00 ± 0.01
0.29 ± 0.12
0.25 ± 0.04
2.15 ± 1.01
2.10 ± 0.02
0.80 ± 0.31
0.73 ± 0.23
1.06 ± 0.48
1.40 ± 0.57
1.40 ± 0.33
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.43
1.40 ± 0.36
1.40 ± 0.52
1.40 ± 0.32
1.40 ± 0.49
1.40 ± 0.37
1.40 ± 0.42
1.40 ± 0.43
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.37
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.38
1.40 ± 0.53
1.40 ± 0.38
1.40 ± 0.37
1.40 ± 0.00
1.40 ± 0.52
1.80 ± 0.55
1.00 ± 0.58
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.40
1.40 ± 0.47
1.60 ± 0.10
1.60 ± 0.10
2.24 ± 0.10
2.16 ± 0.10
2.59 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.52 ± 0.10
2.32 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
4.24 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
2.83 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.70 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.09 ± 0.11
0.09 ± 0.08
0.19 ± 0.03
0.06 ± 0.01
0.06 ± 0.01
0.03 ± 0.04
0.05 ± 0.06
0.22 ± 0.18
0.03 ± 0.03
0.05 ± 0.06
0.10 ± 0.13
0.13 ± 0.10
0.09 ± 0.15
0.11 ± 0.13
0.14 ± 0.02
0.43 ± 0.08
0.08 ± 0.09
0.04 ± 0.02
0.02 ± 0.01
0.03 ± 0.06
0.33 ± 0.16
0.13 ± 0.16
0.05 ± 0.04
0.21 ± 0.17
0.04 ± 0.01
0.03 ± 0.04
0.50 ± 0.17
0.04 ± 0.07
0.04 ± 0.01
0.03 ± 0.05
0.09 ± 0.10
0.03 ± 0.03
32
0.50
0.87
0.33
0.68
0.88
0.84
0.39
0.97
0.49
0.46
0.67
0.55
0.66
0.80
0.38
0.82
0.31
0.37
0.44
0.87
0.87
0.48
0.31
0.50
0.92
0.24
0.65
0.52
0.09
0.20
0.62
0.25
0
0
8
15
27
0
0
0
0
0
0
0
0
0
5
4
0
0
7
0
0
0
0
0
71
0
0
0
5
0
0
0
10
27
8
15
27
11
7
11
11
9
10
8
8
7
5
6
6
13
7
13
11
5
4
5
78
12
6
11
5
4
5
13
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
453707
454078
454100
j3707
18.60
0.15
j4078
17.50
0.15
j4100
20.10
0.15
2002 GP186
K02GI6P
20.30
0.15
2003 KZ18
K03K18Z
21.20
0.15
2003 MT9 K03M09T 18.60
0.15
2006 KL89
K06K89L
18.60
0.15
2006 KL89
K06K89L
18.60
0.15
2006 OF5
K06O05F
19.30
0.15
2006 OF5
K06O05F
19.30
0.15
2006 UR217 K06UL7R 19.80
0.15
2006 UR217 K06UL7R 19.80
0.15
2007 WE55 K07W55E
20.20
0.15
2010 CO1
K10C01O 21.80
0.15
2010 LF86
K10L86F
17.20
0.15
2010 UB8
K10U08B 19.60
0.15
2010 UB8
K10U08B 19.60
0.15
2010 YD3
K10Y03D 20.00
0.15
2011 AM24 K11A24M 20.50
0.15
2011 AM24 K11A24M 20.50
0.15
2011 HJ61
K11H61J
19.30
0.15
2011 JU2
K11J02U 18.40
0.15
2011 OL5
K11O05L
20.20
0.15
2011 OL5
K11O05L
20.20
0.15
2011 VQ5
K11V05Q 20.10
0.15
2011 YB40
K11Y40B 19.10
0.15
2012 OD1
K12O01D 18.60
0.15
2014 JY24
K14J24Y 18.30
0.15
2014 QK434 K14Qh4K 19.10
0.15
2014 TA36
K14T36A 20.70
0.15
2014 US
K14U00S
19.10
0.15
2014 US
K14U00S
19.10
0.15
0.50 ± 0.14
2.96 ± 1.20
0.55 ± 0.01
0.17 ± 0.05
0.47 ± 0.19
0.68 ± 0.22
0.96 ± 0.38
1.07 ± 0.52
0.95 ± 0.55
0.93 ± 0.32
0.89 ± 0.07
0.95 ± 0.40
0.69 ± 0.18
0.29 ± 0.15
2.56 ± 1.22
0.92 ± 0.08
0.88 ± 0.42
0.76 ± 0.35
0.50 ± 0.01
0.51 ± 0.01
1.28 ± 0.57
1.49 ± 0.56
0.28 ± 0.08
0.28 ± 0.11
0.56 ± 0.23
0.42 ± 0.12
0.35 ± 0.09
1.92 ± 1.01
0.30 ± 0.01
0.55 ± 0.23
0.47 ± 0.15
0.56 ± 0.19
1.40 ± 0.39
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.48
1.40 ± 0.48
1.40 ± 0.46
1.40 ± 0.41
1.40 ± 0.53
1.00 ± 0.64
1.00 ± 0.49
1.40 ± 0.11
1.40 ± 0.47
1.40 ± 0.33
1.00 ± 0.67
1.40 ± 0.50
1.16 ± 0.09
1.00 ± 0.61
1.40 ± 0.52
1.40 ± 0.01
1.40 ± 0.01
1.40 ± 0.47
1.40 ± 0.42
1.40 ± 0.42
1.40 ± 0.51
1.40 ± 0.50
1.40 ± 0.42
1.40 ± 0.44
1.40 ± 0.55
1.40 ± 0.00
1.40 ± 0.49
1.40 ± 0.46
1.40 ± 0.42
1.60 ± 0.10
1.60 ± 0.10
1.70 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
6.16 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.47 ± 0.10
0.94 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.83 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.26 ± 0.12
0.02 ± 0.04
0.05 ± 0.01
0.44 ± 0.27
0.03 ± 0.04
0.14 ± 0.17
0.07 ± 0.03
0.06 ± 0.10
0.04 ± 0.03
0.04 ± 0.03
0.03 ± 0.01
0.02 ± 0.05
0.03 ± 0.05
0.04 ± 0.06
0.04 ± 0.05
0.03 ± 0.01
0.03 ± 0.05
0.03 ± 0.04
0.04 ± 0.01
0.04 ± 0.01
0.02 ± 0.04
0.03 ± 0.06
0.19 ± 0.16
0.19 ± 0.16
0.05 ± 0.08
0.22 ± 0.17
0.54 ± 0.26
0.02 ± 0.05
0.46 ± 0.05
0.03 ± 0.03
0.19 ± 0.26
0.13 ± 0.08
33
0.40
0.23
0.26
0.29
0.69
0.30
0.68
0.65
0.63
0.43
0.26
0.44
0.78
0.43
0.49
0.16
0.37
0.54
0.49
0.47
0.97
0.77
1.05
0.60
0.85
0.38
0.36
0.37
0.41
0.69
0.30
0.38
0
0
6
0
0
0
0
0
0
0
8
0
0
0
0
7
0
0
8
14
0
0
0
0
0
0
0
0
5
0
0
0
4
15
6
5
6
4
23
15
46
47
9
6
13
13
5
7
15
12
8
15
13
4
36
13
18
5
5
15
5
11
16
14
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
2014 UV33
K14U33V 17.90
0.15
2014 UF206 K14UK6F
18.80
0.15
2014 UF206 K14UK6F
18.80
0.15
2014 WF365
K14Wa5F
17.20
0.15
2014 XR6
K14X06R 18.30
0.15
2014 XX31
K14X31X 17.50
0.15
2014 YJ14
K14Y14J
18.30
0.15
2014 YS14
K14Y14S
21.10
0.15
2014 YT14
K14Y14T 18.90
0.15
2014 YS34
K14Y34S
20.80
0.15
2014 YB35
K14Y35B 19.00
0.15
2014 YR43
K14Y43R 19.50
0.15
2015 AC17
K15A17C 19.90
0.15
2015 AY245 K15AO5Y 21.20
0.15
2015 AY245 K15AO5Y 21.20
0.15
2015 AK280
K15AS0K 21.80
0.15
2015 BY516
K15Bp6Y 22.30
0.15
2015 CV13
K15C13V 20.30
0.15
2015 DE176 K15DH6E
19.70
0.15
2015 DE176 K15DH6E
19.70
0.15
2015 DX198
K15DJ8X 22.00
0.15
2015 EZ
K15E00Z
20.30
0.15
2015 FZ35
K15F35Z
19.40
0.15
2015 FY117 K15FB7Y 21.30
0.15
2015 FH120
K15FC0H 18.70
0.15
2015 FU332 K15FX2U 17.20
0.15
2015 FD341 K15FY1D 17.70
0.15
2015 FT344 K15FY4T 19.90
0.15
2015 FT344 K15FY4T 19.90
0.15
2015 GY
K15G00Y 21.70
0.15
2015 GK50
K15G50K 20.60
0.15
2015 GK50
K15G50K 20.60
0.15
0.82 ± 0.02
1.52 ± 0.68
1.29 ± 0.06
2.18 ± 0.02
0.86 ± 0.29
1.49 ± 0.53
1.91 ± 0.12
0.30 ± 0.12
1.16 ± 0.47
0.13 ± 0.03
0.28 ± 0.01
0.37 ± 0.13
0.67 ± 0.28
0.37 ± 0.03
0.39 ± 0.18
0.36 ± 0.12
0.24 ± 0.12
0.44 ± 0.13
0.68 ± 0.04
0.57 ± 0.29
0.35 ± 0.10
0.19 ± 0.05
0.64 ± 0.21
0.38 ± 0.17
0.75 ± 0.26
0.94 ± 0.36
1.25 ± 0.48
0.75 ± 0.24
0.76 ± 0.21
0.14 ± 0.05
0.61 ± 0.30
0.46 ± 0.03
1.40 ± 0.00
1.00 ± 0.53
1.12 ± 0.05
1.40 ± 0.00
1.40 ± 0.45
1.40 ± 0.39
1.12 ± 0.06
1.40 ± 0.51
1.00 ± 0.52
1.40 ± 0.41
1.40 ± 0.00
1.40 ± 0.47
1.00 ± 0.57
1.40 ± 0.12
1.40 ± 0.59
1.40 ± 0.41
1.00 ± 0.64
1.40 ± 0.35
1.40 ± 0.07
1.40 ± 0.59
1.40 ± 0.38
1.40 ± 0.39
1.40 ± 0.41
1.00 ± 0.60
1.40 ± 0.46
1.40 ± 0.53
1.40 ± 0.43
1.40 ± 0.39
1.40 ± 0.34
1.40 ± 0.50
1.00 ± 0.62
0.99 ± 0.06
3.29 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.39 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.31 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
8.64 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.21 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.18 ± 0.03
0.02 ± 0.02
0.03 ± 0.01
0.05 ± 0.01
0.11 ± 0.17
0.08 ± 0.06
0.02 ± 0.01
0.07 ± 0.09
0.04 ± 0.02
0.50 ± 0.23
0.57 ± 0.07
0.20 ± 0.13
0.04 ± 0.04
0.04 ± 0.02
0.04 ± 0.09
0.03 ± 0.04
0.04 ± 0.03
0.07 ± 0.04
0.05 ± 0.01
0.07 ± 0.11
0.02 ± 0.02
0.36 ± 0.19
0.08 ± 0.11
0.04 ± 0.04
0.11 ± 0.11
0.26 ± 0.27
0.09 ± 0.08
0.03 ± 0.03
0.03 ± 0.04
0.18 ± 0.19
0.03 ± 0.03
0.05 ± 0.01
34
1.58
0.68
0.16
0.20
0.56
0.67
0.12
0.57
0.95
0.46
0.52
0.36
0.67
0.43
2.84
0.36
0.55
0.69
0.19
0.25
1.31
0.62
0.57
0.62
0.25
0.47
0.77
0.60
0.46
0.24
0.48
0.27
20
0
23
7
0
0
7
0
0
0
6
0
0
13
0
0
0
0
9
0
0
0
0
0
0
0
0
0
0
0
0
24
46
23
7
10
6
7
5
18
5
6
9
34
13
60
4
8
8
9
4
9
10
5
43
11
9
4
4
7
4
9
16
16
Table 4-Continued
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
2015 GN50
K15G50N 20.20
0.15
2015 HF11
K15H11F
19.40
0.15
2015 JF11
K15J11F
21.20
0.15
2015 KH157
K15KF7H 20.00
0.15
2015 KL157
K15KF7L
19.10
0.15
2015 KL157
K15KF7L
19.10
0.15
2015 LK24
K15L24K 21.60
0.15
2015 MQ130 K15MD0Q 20.90
0.15
2015 MQ130 K15MD0Q 20.90
0.15
2015 NA14
K15N14A 22.00
0.15
2015 OA22
K15O22A 20.00
0.15
2015 OS35
K15O35S
19.10
0.15
2015 OS35
K15O35S
19.10
0.15
2015 PD
K15P00D 19.30
0.15
2015 PM57
K15P57M 18.60
0.15
2015 QM3
K15Q03M 20.30
0.15
2015 RS83
K15R83S
19.40
0.15
2015 RR150
K15RF0R 19.70
0.15
2015 SF20
K15S20F
19.70
0.15
2015 SS20
K15S20S
22.40
0.15
2015 TK237
K15TN7K 22.60
0.15
2015 TW346 K15TY6W 18.60
0.15
2015 UK52
K15U52K 20.10
0.15
2015 VZ145
K15VE5Z
23.70
0.15
2015 WM16 K15W16M 21.80
0.15
2015 XB130
K15XD0B 21.80
0.15
2015 XY378
K15Xb8Y 19.60
0.15
0.29 ± 0.11
1.11 ± 0.44
0.17 ± 0.05
0.58 ± 0.23
1.45 ± 0.31
0.36 ± 0.01
0.31 ± 0.11
0.36 ± 0.07
0.55 ± 0.26
0.09 ± 0.02
0.79 ± 0.34
1.26 ± 0.01
1.42 ± 0.01
0.62 ± 0.22
0.59 ± 0.21
0.27 ± 0.05
0.47 ± 0.13
0.34 ± 0.12
0.40 ± 0.16
0.26 ± 0.10
0.23 ± 0.07
1.26 ± 0.46
0.21 ± 0.06
0.16 ± 0.06
0.39 ± 0.06
0.33 ± 0.13
0.31 ± 0.11
0.18 ± 0.12
0.02 ± 0.01
0.20 ± 0.11
0.05 ± 0.11
0.02 ± 0.01
0.30 ± 0.05
0.04 ± 0.07
0.06 ± 0.03
0.03 ± 0.03
0.34 ± 0.18
0.03 ± 0.03
0.03 ± 0.00
0.02 ± 0.00
0.09 ± 0.10
0.19 ± 0.22
0.19 ± 0.15
0.14 ± 0.13
0.20 ± 0.19
0.15 ± 0.14
0.03 ± 0.03
0.03 ± 0.04
0.04 ± 0.04
0.35 ± 0.22
0.02 ± 0.05
0.02 ± 0.01
0.03 ± 0.04
0.26 ± 0.23
1.40 ± 0.46
1.40 ± 0.42
1.40 ± 0.40
1.40 ± 0.49
1.00 ± 0.38
0.40 ± 0.00
1.40 ± 0.44
0.54 ± 0.10
1.00 ± 0.61
1.40 ± 0.32
1.00 ± 0.56
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.49
1.40 ± 0.27
1.40 ± 0.37
1.40 ± 0.46
1.40 ± 0.53
1.40 ± 0.46
1.40 ± 0.40
1.40 ± 0.41
1.40 ± 0.45
1.40 ± 0.52
1.10 ± 0.16
1.40 ± 0.48
1.40 ± 0.46
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.50 ± 0.10
1.50 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
5.95 ± 0.10
6.34 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
1.60 ± 0.10
0.36
0.82
0.34
0.22
0.76
1.05
0.39
0.36
0.83
0.45
0.11
0.17
0.15
0.57
0.22
0.61
0.52
0.60
0.52
0.44
0.53
0.44
0.34
1.36
0.17
0.58
0.32
0
0
0
0
0
9
0
5
0
0
0
20
10
0
0
0
0
0
0
0
0
0
0
0
8
0
0
5
10
6
9
48
12
8
7
4
5
4
22
10
8
7
9
5
13
8
7
7
15
5
5
8
9
5
35
Table 5. Measured diameters (d) and albedos (pV ) of non-NEA asteroids observed during
Year 2 of the NEOWISE Reactivation mission. Asteroids may be identified by numbers,
provisional designations, or via the MPC packed format. Beaming η, H, G, the amplitude
of the 4.6 µm light curve (W2 amp., in mag), and the numbers of observations used in the
3.4 µm (nW 1) and 4.6 µm (nW 2) wavelengths are also reported. Only the first 10 lines are
shown; the remainder are available in electronic format through the journal website.
Object
Packed
H
G
d (km)
pV
η
pIR/pV
W2 amp.
nW 1
nW 2
10
13
13
19
19
21
21
23
30
33
00010
5.46
0.12
00013
6.77
0.12
00013
6.77
0.12
00019
7.20
0.12
00019
7.20
0.12
00021
7.45
0.24
00021
7.45
0.24
00023
7.09
0.24
00030
7.67
0.24
00033
8.60
0.24
450.53 ± 200.23
207.98 ± 46.68
192.79 ± 53.36
176.97 ± 56.71
182.71 ± 40.61
102.07 ± 24.56
99.71 ± 22.62
93.99 ± 20.14
105.70 ± 23.25
54.39 ± 11.84
0.05 ± 0.05
0.08 ± 0.04
0.09 ± 0.06
0.06 ± 0.07
0.06 ± 0.03
0.18 ± 0.08
0.18 ± 0.05
0.29 ± 0.14
0.19 ± 0.11
0.23 ± 0.13
0.95 ± 0.23
1.00 ± 0.35
1.00 ± 0.39
0.95 ± 0.21
0.95 ± 0.14
1.00 ± 0.39
1.00 ± 0.38
0.95 ± 0.20
0.95 ± 0.19
0.95 ± 0.19
1.50 ± 0.10
1.00 ± 0.60
1.00 ± 0.60
1.50 ± 0.10
1.50 ± 0.10
1.00 ± 0.60
1.00 ± 0.60
1.50 ± 0.10
1.50 ± 0.10
1.50 ± 0.10
0.05
0.16
0.32
0.25
0.32
0.50
0.35
0.23
0.23
0.36
4
5
9
10
8
11
12
9
6
7
4
5
9
10
8
11
13
9
7
7
Table 6. Measured diameters and albedos for objects observed during the Year 2
Reactivation mission that may be accessible by spacecraft, following the NHATS criteria.
Also listed is the minimum mission round trip time in days for each object from Barbee
et al. (2013).
Number
Designation
D (km)
pV
Minimum round trip (days)
(35107)
1991 VH
(163899)
2003 SD220
(363505)
2003 UC20
(424392)
2007 YJ
2011 AM24
2015 GY
2015 NA14
0.91 ± 0.03
0.80 ± 0.02
1.88 ± 0.01
0.24 ± 0.10
0.50 ± 0.01
0.14 ± 0.05
0.09 ± 0.02
0.33 ± 0.04
0.31 ± 0.04
0.03 ± 0.00
0.05 ± 0.06
0.04 ± 0.01
0.18 ± 0.19
0.34 ± 0.18
354
122
290
98
130
346
170
36
|
1110.2426 | 1 | 1110 | 2011-10-11T16:47:22 | Compositions of Hot Super-Earth Atmospheres: exploring Kepler Candidates | [
"astro-ph.EP",
"astro-ph.SR"
] | This paper outlines a simple approach to evaluate the atmospheric composition of hot rocky planets by assuming different types of planetary composition and using corresponding model calculations. To explore hot atmospheres above 1000 K, we model the vaporization of silicate magma and estimate the range of atmospheric compositions according to the planet's radius and semi-major axis for the Kepler February 2011 data release. Our results show 5 atmospheric types for hot, rocky super-Earth atmospheres, strongly dependent on the initial composition and the planet's distance to the star. We provide a simple set of parameters that can be used to evaluate atmospheric compositions for current and future candidates provided by the Kepler mission and other searches. | astro-ph.EP | astro-ph | Draft version September 11, 2018
Preprint typeset using LATEX style emulateapj v. 11/10/09
1
1
0
2
t
c
O
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
2
4
2
.
0
1
1
1
:
v
i
X
r
a
COMPOSITIONS OF HOT SUPER-EARTH ATMOSPHERES: EXPLORING KEPLER CANDIDATES
Max Planck Institut fuer Astronomie, Koenigstuhl 17, 69117, Heidelberg, Germany
Y. Miguel
Max Planck Institut fuer Astronomie, Koenigstuhl 17, 69117, Heidelberg, Germany
Harvard Smithsonian Center for Astrophysics, 60 Garden St., 02138 MA,Cambridge, USA
L. Kaltenegger
Planetary Chemistry Laboratory, Dpt. of Earth & Planetary Sciences and Mc Donnell Center for the Space Sciences, Washington
University, St. Louis, MO 63130, USA
B. Fegley Jr.
Harvard Smithsonian Center for Astrophysics, 60 Garden St., 02138 MA,Cambridge, USA
L. Schaefer
Draft version September 11, 2018
ABSTRACT
This paper outlines a simple approach to evaluate the atmospheric composition of hot rocky planets
by assuming different types of planetary composition and using corresponding model calculations. To
explore hot atmospheres above 1000 K, we model the vaporization of silicate magma and estimate
the range of atmospheric compositions according to the planet's radius and semi-major axis for the
Kepler February 2011 data release. Our results show 5 atmospheric types for hot, rocky super-Earth
atmospheres, strongly dependent on the initial composition and the planet's distance to the star. We
provide a simple set of parameters that can be used to evaluate atmospheric compositions for current
and future candidates provided by the Kepler mission and other searches.
Subject headings: Astrobiology - Atmospheric effects - Planets and satellites: atmospheres - Planets
and satellites: composition
1. INTRODUCTION
In recent years the search for extrasolar planets has
resulted in the discovery of the first confirmed rocky
planets with minimum masses below 10 M⊕ (hereafter
super-Earths) (Udry et al. 2007; Leger et al. 2009) and
radii below 2 R⊕, consistent with rocky planet mod-
els (Borucki et al. 2011; Batalha et al. 2011). Many
of the detected super-Earths are located extremely close
to their stars, at distances much less than Mercury to the
Sun. These close-in hot super-Earths can reach temper-
atures high enough to melt the silicates present on their
surfaces by irradiation from their host star alone. This
causes outgassing and the subsequent formation of the
planet's atmosphere.
The formation of a rocky planet's atmosphere is a com-
plex process built up from initial capture and degassing
of material of meteoritic composition. Our model is
based on work by Schaefer & Fegley (2004, 2009) and
assumes, as a first order approximation, that at temper-
atures Tp >1000 K, the material present on the surface of
the planet vaporizes and forms an outgassed atmosphere.
In this process, the atmosphere is strongly dependent
on the planetary composition (Schaefer & Fegley 2009;
Valencia et al 2010). Schaefer & Fegley (2009) mod-
eled the specific case of CoRoT-7b and showed that these
close-in planets may have a silicate atmosphere as a re-
sult of outgassing of magma on the surface.
[email protected]
The Kepler mission recently announced 1235 planetary
candidates (Borucki et al. 2011), of which 615 are con-
sistent with models of rocky interior (R < 2± 0.5 R⊕,
accounting for the 25% error in radii in the Kepler data
set). 193 of these Kepler objects of interest (KOI) have
potentially extremely high temperatures on their surfaces
(Tp >1000 K, when assuming a 0.01 albedo), due to the
intense stellar irradiation to which they are exposed, be-
cause of their proximity to the central star.
In this work, we explore the composition of initial
planetary atmospheres of hot super-Earths in the Kepler
planet candidate sample, according to their semi major
axis. Our approach can be applied to current and future
candidates provided by the Kepler mission and other ex-
oplanet searches.
2. OUTGASSING MODEL
We use the MAGMA code (Fegley & Cameron 1987;
Schaefer & Fegley 2004) which calculates the equilib-
rium between the melt and vapor in a magma exposed
at temperatures higher than 1000 K (Hastie et al. 1982,
1984; Hastie & Bonnell 1985, 1986), for Al, Ca, Fe, K,
Mg, Na, O, Si, Ti and their compounds.
We assume that the initial composition of a hot super-
Earth is similar to the composition of komatiites, which
are ultramafic lavas that were erupted during the Ar-
chaean on Earth (3.8 to 2.5 billion years ago), when
Earth had a higher surface temperature. This choice
is also supported by the work by Rubie et al.
(2004),
who showed, for the case of the Earth and Mars, that a
Rp>10 R⊕
5<Rp<10 R⊕
2.5<Rp<5 R⊕
2<Rp<2.5 R⊕
1.5<Rp<2 R⊕
Rp<1.5 R⊕
2
)
K
(
e
r
u
t
a
r
e
p
m
e
T
e
v
i
t
c
e
f
f
E
r
a
l
l
e
t
S
9000
8000
7000
6000
5000
4000
3000
0.01
0.1
Semimajor axis (AU)
1
Fig. 1. -- Effective stellar temperature vs semi-major axis of all
Kepler planetary candidates, in different point sizes, according to
their radii. Planet candidates with Tp >1000 K and Rp ≤ 2.5 R⊕,
are shown in red. The cooler planets (in the same radius range)
are shown in black. The planet candidates with Rp >2.5 R⊕ are
shown in grey.
more massive planet has a lower composition of FeO in
the mantle than a less massive one, a consequence of the
solubility of oxygen in liquid iron-alloy, which increases
with temperature. Table 1 shows komatiite and bulk sil-
icate Earth (BSE) compositions (see section 3.1.1) for
comparison.
3. ATMOSPHERIC COMPOSITION OF KEPLER
CANDIDATES
The planet's equilibrium temperature due to irradia-
tion by a central star of radius R⋆ and effective temper-
ature Tef f,⋆, is given by equation 1,
Tp = Tef f,⋆(cid:18) (1 − A)R2
4a2
⋆
1
4
(cid:19)
(1)
the planet's
where a is
semi-major axis and A
its bond albedo. We adopt an albedo of 0.01
(Miller-Ricci & Fortney 2010). Different values for the
albedo can change our results slightly (see discussion).
Here we set Tp equal to the planet's initial surface tem-
perature, to explore the resulting atmospheric composi-
tion.
Figure 1 shows the stellar effective temperature ver-
sus planet semi-major axis for all the Kepler candidates,
where different point sizes indicate different planet ra-
dius. Using equation 1, we separate the planets accord-
ing to their radius and surface temperature. The planet
candidates with derived Tp > 1000 K and Rp ≤ 2.5 R⊕,
are shown in red and are the sample we focus on here.
The cooler planets (in the same radius range) are shown
in black. The planet candidates with Rp >2.5 R⊕ are
shown in grey.
3.1. Simulation Results
We perform outgassing simulations for temperatures
between 1000 and 3500 K, and apply our results to
planet candidates in the February 2011 Kepler release
with Rp ≤ 2.5 R⊕ and Tp > 1000 K. Table 2, lists their
semimajor axis, radii, surface temperatures (obtained
adopting an albedo of 0.01), stellar effective tempera-
tures and stellar radius.
(a)
(b)
Fig. 2. -- Tef f,⋆ vs. a for the Kepler candidates analyzed in this
work (upper panel) . KOIs with radius between 2-2.5 R⊕ (large
points), radius between 1.5-2 R⊕ (medium points) and Rp <1.5 R⊕
(small points). Lines of constant temperatures are shown. Planet
surface temperatures vs. partial pressures of the gases vaporized
from a komatiite magma (lower panel). Temperatures at which the
dominant gases change are indicated.
Following Schaefer & Fegley (2009), we explore the
partial pressures of all the gases vaporized at differ-
ent temperatures, assuming atmospheres free of volatile
gases such as H, C, N, S (see discussion). We model the
initial planetary atmosphere characteristics depending
on the radius and semi-major axis of the Kepler candi-
dates. Figure 2(a) shows the planet candidates that orbit
their star with short periods and accordingly high tem-
peratures. The candidates are shown in different point
sizes according to their radius: 2-2.5 R⊕ (large points),
1.5-2 R⊕ (medium points) and candidates with Rp <1.5
R⊕ (small points). The lines indicate constant planet
surface temperatures. The blue, violet, light blue, grey
and black line in figure 2(a) indicate Tp of 2096, 2460,
2735, 2974 and 3170 K, respectively. These temperatures
indicate significant changes in the partial pressures of the
dominant gases. Figure 2(b) shows the partial pressures
of the gases as a function of the planet surface tempera-
ture.
Planets located below the blue line in figure 2(a) have
calculated surface temperatures Tp <2096 K. As seen
in figure 2(b), these planets are characterized by par-
tial pressures dominated by Na, O2, O and Fe (from the
Compositions adopted for the magma
TABLE 1
Oxide
SiO2 MgO Al2O3 TiO2 Fe2O3 FeO CaO Na2O K2O
Komatiite (%)
Bulk Silicate Earth (%)
47.10
45.97
29.60
36.66
4.04
4.77
0.24
0.18
12.80
0.0
0.0
8.24
5.44
3.78
0.46
0.35
0.09
0.04
3
Planetary and stellar parameters for Kepler candidates
with radii less than 2.5 R⊕ and temperatures higher than
1000 K.
TABLE 2
KOI
a (AU) Rp (R⊕)
Tp (K) a Tef f,⋆ (K) R⋆ (R⊙)
69.01
70.02
72.01
85.02
85.03
107.01
112.02
115.02
117.02
117.03
123.01
124.01
137.03
139.02
0.056
0.046
0.018
0.035
0.084
0.075
0.048
0.076
0.058
0.043
0.071
0.111
0.046
0.045
1.6
1.6
1.3
1.7
2.
2.1
1.7
2.2
1.3
1.3
2.3
2.3
2.3
1.2
1129
1001
1946
1987
1282
1025
1414
1251
1142
1326
1188
1006
1335
1272
5480
5342
5491
6006
6006
5816
5839
6202
5725
5725
5897
6076
5289
5921
1.03
0.7
0.98
1.66
1.66
1.01
1.22
1.34
1
1
1.25
1.32
1.27
0.9
Note. -- Table 2 is published in its entirety in the electronic
edition. A portion is shown here for guidance regarding its form
and content.
a Calculated adopting an albedo of 0.01
highest to the lowest partial pressures). In the same way,
planets located between the blue and the violet line have
temperatures 2096 < Tp < 2460 K. In this case, the
partial pressure of SiO becomes higher than Fe. With
surface temperatures between 2460 and 2735, the planets
are located between the violet and the light-blue line. For
these surface temperatures, the partial pressure of SiO is
higher than O. When the surface temperature is between
2735 < Tp < 2974 K (between the light-blue and the
grey line), the partial pressure of SiO is higher than O2
and for 2974 < Tp < 3170 K (between the grey and the
black line) SiO becomes more abundant than monatomic
Na. Finally a planet with Tp > 3170 K is located above
the black line and is characterized by the highest partial
pressure of SiO, followed by O2, monatomic O and Na.
3.1.1. Different types of Atmospheres
In order to link the partial pressures to the result-
ing atmospheric composition, we calculate the column
density (σi) of each gas i, as a function of the gas'
partial pressure, Pi,
for planetary masses between 1
and 10 M⊕, corresponding to radii between ∼ 1-2.5
R⊕ (Valencia, Sasselov & O'Connell 2007; Seager et al
2007; Sotin et al. 2007) using equation 2.
σi =
PiNA
gµi
(2)
where NA is Avogadro's number, µi is the molecular
weight of the species and g is the gravitational surface
acceleration of each planet. We find that the composi-
tions of the atmospheres do not show significant changes
when considering different planetary masses. This means
that the surface temperature and therefore the distance
to the host star is the dominating factor in determining
the composition of the atmosphere, assuming a certain
planetary composition.
We adopt a komatiite composition for hot super-Earths
and explore the effects of initial composition on the re-
sults by comparing with a Bulk Silicate Earth (B.S.E)
initial composition in Figure 3. Figure 3(a) (komati-
ite) and 3(b) (BSE composition) show the stellar effec-
tive temperature vs. the planet's distance to the star.
Lines show the temperatures that separate the different
classes of planetary atmospheres. Areas highlighted in
gray scale indicate the regions of different types of atmo-
spheres. The planets in each one of these areas are char-
acterized by different types of atmospheres as indicated
below. Figures 3(c) and 3(d) show the column density of
all the gases vaporized vs. the planet surface tempera-
ture for a 10 M⊕ planet with komatiite and BSE as the
initial composition, respectively. Since we find similar
results for planets with different masses, the column den-
sities shown in Figures 3(c) and 3(d) are representative
of planets with masses between 1-10 M⊕. The dominant
gases are Fe, Mg, Na, O, O2 and SiO.
Our results divide hot rocky planet atmospheres into 5
classes. Temperatures that limit each class change when
considering different magma compositions. For komatiite
composition:
Atmospheres Type I (Tp < 2033 K), characterized
by the presence of monatomic Na, O2, monatomic O and
monatomic Fe in order of abundance.
Atmospheres Type II (2033 < Tp < 2588 K),
SiO becomes more abundant than monatomic Fe. The
atmosphere is characterized by monatomic Na, O2,
monatomic O, SiO, monatomic Fe and monatomic Mg
from the most to the least abundant.
Atmospheres Type III (2588 < Tp < 2890 K),
monatomic Mg becomes more abundant than monatomic
Fe. The gases with the highest column densities are
monatomic Na, O2, monatomic O, SiO, monatomic Mg
and monatomic Fe. Note that the column densities of
O2, monatomic O and SiO are almost equal in this tem-
perature range.
Atmospheres Type VI (2890 < Tp < 3168 K), SiO
becomes more abundant than O2 and the atmosphere
becomes silicate and monatomic Na dominated.
Atmospheres Type V (Tp > 3168 K), dominated by
silicate oxide, followed closely by monatomic O, O2 and
monatomic Na.
Of the 193 Kepler planetary candidates studied in this
work (see Figure 3(a)), 189 are characterized by Type I
atmospheres when adopting komatiite as the initial com-
position. Three planets are characterized by Type II (be-
tween the black and the dark-red line in figure 3(a)), no
planets by Type III (between the dark-red and red lines),
only one by Type IV (located between the red and yel-
4
(a)
(b)
(c)
(d)
Fig. 3. -- Stellar effective temperature vs. semi-major axis of Kepler planetary candidates (upper panels, see figure 2(a) for detailed
explanation) and column densities of the atmospheric gases vs. planet surface temperature (lower panels). The column densities were
calculated for a planet of 10 M⊕, but are representative of planets with masses between 1-10 M⊕ (see section 3.1.1). We adopt two different
planetary compositions: komatiite (figures 3(a), 3(c)) and bulk silicate of the Earth (figures 3(b), 3(d)).
low lines) and none by Type V atmospheres (above the
yellow line) , if we only consider the irradiation from the
star for heating (see discussion).
When adopting the BSE composition, the temperature
limits between the 5 classes change slightly to 2070, 2120,
2996 and 3278 K, respectively. For a BSE composition,
190 planetary candidates are characterized by Type I at-
mospheres (located below the black line in figure 3(b)),
only 1 planet by Type II atmosphere (between the black
and the dark-red lines), 1 by Type III (between the dark-
red and red lines), 1 by Type IV (located between the red
and yellow lines) and none by Type V (above the yellow
line). As a result, note that, for this planetary sam-
ple, the initial composition is not important when the
planet's temperature is high for atmospheres of Type IV
and V. Nevertheless, for lower temperatures (Types I, II
and III) it becomes an important factor, since a small
difference in temperature can change the dominant gases
in the planetary atmosphere.
4. DISCUSSION: THE SCOPE OF OUR WORK
The composition of the atmosphere depends strongly
on the composition of the planetary interior. We show
the changes in the characterization when adopting two
known compositions taken from our Solar system, but
substantially different compositions are possible.
Following Schaefer & Fegley (2009), we consider at-
mospheres produced by the outgassing of a volatile-
free magma. At these extreme temperatures, elements
such as H, C, N and S should escape from the at-
mosphere in short timescales and consequently are not
present in the atmosphere. This assumption is sup-
ported by observations of volatile loss in rocky bodies
(Spencer & Schneider 1996; Fegley 2004) and studies in
hot-Jupiters (Baraffe et al. 2005) and hot super-Earths
(Leitzingers et al. 2011). If the volatiles were not lost,
the atmosphere would be dominated by volatiles.
In
this paper we only concentrate on close-in, hot plan-
etary candidates, consistent with volatile free atmo-
spheres. Cooler planets that could potentially provide
habitable conditions are not discussed here (see e.g.
Kaltenegger & Sasselov (2011)).
In this work we have adopted a bond albedo for the
planets of A=0.01. This assumption is supported by
observations (Rogers et al.
2006,
(Sudarsky, Burrows & Pinto
2008)
2000;
2008;
Miller-Ricci & Fortney 2010) of hot exoplanets, which
Burrows, Ibgui & Hubeny
and modeling
2009; Rowe et al.
find very low albedo. Although there are currently no
indications that suggest higher albedos for hot rocky
planets, we studied the effect of a higher albedo on our
results. When changing the albedo from 0.01 to e.g.
0.2, the size of the sample is affected. For an albedo of
0.2, only 146 planetary candidates with radius less than
2.5 R⊕ have Tp > 1000 K. Of these 146 planets, 144 are
characterized by atmospheres Type I, 1 by Type II, 1
by Type III, 0 by Type IV and 0 by Type V (adopting
komatiite composition).
We assume here that the planets are only heated by the
irradiation of the star. Other heat sources like short-lived
radioisotopes, impacts, core formation and tidal effects
can increase the surface temperature. Henning et al.
(2009) showed that the insolation is the most important
heating source for planets with periods larger than ∼ 2
days and for planets with shorter periods and non ec-
centric orbits. For very eccentric planets (e >0.1) which
orbit their star with periods less than ∼ 2 days, tidal
heating dominates over insolation. We explore the ef-
fects of tidal heating in the Kepler planetary candidates
with P < 2 days, using eq. 3 (Henning et al. 2009;
Murray & Dermott 2005),
Tp,tidal+ins = (cid:18) πR2
p(1 − A) L⋆
4πa2 + 21
2
pσ
4πR2
GR5
pne2
a6 M 2
⋆
k2
Q
1
4
(cid:19)
(3)
where, k2 is the second-order Love number of the planet.
For a super-Earth we adopt 0.3 (Henning personal com-
munication), Q is the quality factor of the planet, n its
mean motion and e the eccentricity, which is unknown
for the Kepler candidates. We adopt different values for
e and Q. Most of the candidates that might be affected
are located very close to the star (a less than ∼ 0.03
AU), where we expect almost circular orbits. We adopt
e =0.001 and e =0.01 and two different values for Q: 200
and 20000, which represent planets with limited partial
melting or significant partial melting in the mantle, re-
spectively (Henning et al. 2009), to explore this effect.
In the four cases analyzed, we find that there is no signif-
icant change in the surface temperatures of these planets
when considering heating by insolation plus tidal effects
compared to heating by insolation only. The role of tides
5
in these planets, if occurring, will be to increase the con-
vective vigor of the mantle, causing a larger fraction of
total mantle volume to be accessible to surface degassing,
but will not change the surface temperature significantly.
5. CONCLUSIONS
In this work we explore the different atmospheric com-
position of hot super-Earth candidates from the February
Kepler database according to their radii, semi-major axis
and stellar effective temperature. 615 Kepler planet can-
didates have radii less than 2.5 R⊕, of which 193 have
calculated surface temperatures larger than 1000 K (as-
suming an albedo of 0.01), and are the focus of this work.
We model the initial planetary atmospheric compositions
for komatiite and bulk silicate Earth magma composi-
tions. The planet surface temperatures were calculated
assuming insolation only. We discuss the consequences of
adopting different sources of heating and find that tidal
heating is not important when adopting low eccentrici-
ties, consistent with the Kepler sample.
The results indicate 5 types of atmospheres, based on
the abundance of the dominant gases (Fe, Mg, Na, O,
O2 and SiO). When adopting komatiite composition, 189
of 193 Kepler candidates present a Type I model atmo-
sphere dominated by monatomic Na, O2, monatomic O
and monatomic Fe. Three planetary candidates are char-
acterized by Type II atmospheres, where SiO becomes
more abundant than monatomic Fe, no candidates with
Type III atmospheres, where monatomic Mg is more
abundant than monatomic Fe, and only one planet is
characterized by Type IV, with a large abundance of SiO.
Finally, there is no Kepler planetary candidate with an
atmosphere of Type V, which is completely dominated
by SiO.
This atmospheric characterization is almost indepen-
dent of the planetary mass between 1-10 M⊕ and low
eccentricities. The initial magma composition is impor-
tant for the resulting atmospheric gases, especially for
lower surface temperatures. We find that the distance
from the host star is the dominant factor to characterize
a rocky planet's atmosphere.
We are grateful to Dimitar Sasselov and Wade Henning
for stimulating discussions.
REFERENCES
Baraffe I., et al., 2005, A&A, 436, L47
Batalha N. et al., 2011, ApJ, 729, 27
Borucki et al., 2011, ApJ, 736, 19
Burrows A., Ibgui L., Hubeny I., 2008, ApJ, 682, 1277
Fegley B. Jr., 2004, in Meteorites, Comets and Planets, ed. A.M.
Davis, Vol 1 Treatise on Geochemistry ed. K.K. Turekian & H.
D., Holland (Boston, MA Elsevier) 487
Fegley B. Jr. & Cameron A. G. W., 1987, Earth and Planet. Sci.
Letters, 82, 207
Hastie J. W. et al., 1982, in High Temp.-High Press., 14,669
Hastie J. W. et al., 1984, in Ribeiro da Silva, M. A. V. (Ed.),
Thermochemistry and its Applications to Chemical and
Biochemical Systems, Reidel, Dordrecht, 235
Hastie J. W. & Bonnell D. W., 1985, Sci. , 19,275
Hastie J. W. & Bonnell D. W., 1986, J. Non-Crystalline Solids,
84, 151
Leitzingers M., et al., 2011, Planetary and Space Science, 59, 1472
Miller-Ricci E., Seager S. & Sasselov D., 2009, ApJ, 690, 1056
Miller-Ricci E. & Fortney J. J., 2010, ApJ, 716, 74
Murray C. D. & Dermott S. F., 2005, Solar Systema Dynamics
(New York: Cambridge Univ. Press)
Rogers J. C., Apai D., Lpez-Morales M., Sing D. K., Burrows A.,
2009, ApJ, 707, 1707
Rowe J. F. et al., 2006, ApJ, 646, 1241
Rowe J. F. et al., 2008, ApJ, 689, 1345
Rubie D. C., Gessmann C. K. & Frost D. J., 2004, Nature, 429, 58
Schaefer, L., & Fegley, Jr., B. 2004, Icarus, 169, 216
Schaefer L. & Fegley B. Jr., 2009,ApJ, 703, 113
Seager S., Kuchner M., Hier-Majumder C. A., Militzer B., 2007,
ApJ, 669, 1279
Sotin C., Grasset O., Mocquet A., 2007, Icarus, 191, 337
Spencer J. R. & Schneider N. M., 1996, Ann Rev. Earth Planet.
Henning W. G., Connell R. J. O. & Sasselov D. D., 2009, ApJ,
Sci., 24, 125
707, 1000
Kaltenegger L. & Sasselov D.D., 2011, ApJ, 736, 2
Leger et al., 2009, A&A, 506, 287
Sudarsky D., Burrows A., Pinto P., 2000, ApJ, 538, 885
Udry et al., 2007, A&A, 469, 43
Valencia D., Sasselov D. D., O' Connell R., 2007, ApJ, 665, 1413
6
Valencia D., Ikoma M., Guillot T., Nettelmann N., 2010, A&A,
516, 20
|
1109.2949 | 1 | 1109 | 2011-09-13T23:25:58 | Young Solar System's Fifth Giant Planet? | [
"astro-ph.EP"
] | Recent studies of solar system formation suggest that the solar system's giant planets formed and migrated in the protoplanetary disk to reach resonant orbits with all planets inside 15 AU from the Sun. After the gas disk's dispersal, Uranus and Neptune were likely scattered by gas giants, and approached their current orbits while dispersing the transplanetary disk of planetesimals, whose remains survived to this time in the region known as the Kuiper belt. Here we performed N-body integrations of the scattering phase between giant planets in an attempt to determine which initial states are plausible. We found that the dynamical simulations starting with a resonant system of four giant planets have a low success rate in matching the present orbits of giant planets, and various other constraints (e.g., survival of the terrestrial planets). The dynamical evolution is typically too violent, if Jupiter and Saturn start in the 3:2 resonance, and leads to final systems with fewer than four planets. Several initial states stand out in that they show a relatively large likelihood of success in matching the constraints. Some of the statistically best results were obtained when assuming that the solar system initially had five giant planets and one ice giant, with the mass comparable to that of Uranus and Neptune, was ejected to interstellar space by Jupiter. This possibility appears to be conceivable in view of the recent discovery of a large number free-floating planets in interstellar space, which indicates that planet ejection should be common. | astro-ph.EP | astro-ph | Young Solar System’s Fifth Giant Planet?
David Nesvorn´y
Department of Space Studies, Southwest Research Institute, 1050 Walnut St., Suite 300,
Boulder, Colorado 80302, USA
ABSTRACT
Recent studies of solar system formation suggest that the solar system’s giant
planets formed and migrated in the protoplanetary disk to reach resonant orbits
with all planets inside ∼15 AU from the Sun. After the gas disk’s dispersal,
Uranus and Neptune were likely scattered by gas giants, and approached their
current orbits while dispersing the transplanetary disk of planetesimals, whose
remains survived to this time in the region known as the Kuiper belt. Here we
performed N-body integrations of the scattering phase between giant planets in
an attempt to determine which initial states are plausible. We found that the
dynamical simulations starting with a resonant system of four giant planets have a
low success rate in matching the present orbits of giant planets, and various other
constraints (e.g., survival of the terrestrial planets). The dynamical evolution is
typically too violent, if Jupiter and Saturn start in the 3:2 resonance, and leads
to final systems with fewer than four planets. Several initial states stand out in
that they show a relatively large likelihood of success in matching the constraints.
Some of the statistically best results were obtained when assuming that the solar
system initially had five giant planets and one ice giant, with the mass comparable
to that of Uranus and Neptune, was ejected to interstellar space by Jupiter. This
possibility appears to be conceivable in view of the recent discovery of a large
number free-floating planets in interstellar space, which indicates that planet
ejection should be common.
1
1
0
2
p
e
S
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
4
9
2
.
9
0
1
1
:
v
i
X
r
a
1.
Introduction
Studies of giant planets’ interaction with a protoplanetary gas disk show that their
orbits radially migrate, and typically achieve a compact configuration, in which pairs of the
neighbor planets are locked in the mean motion resonances (Kley 2000, Masset & Snellgrove
2001). The resonant planetary systems emerging from protoplanetary disks can become
dynamically unstable after the gas disappears, leading to a phase when planets scatter off
– 2 –
of each other. This model can explain the observed resonant exoplanets, commonly large
exoplanet eccentricities (Weidenschilling & Marzari 1996), and microlensing data that show
evidence for a large number of planets that are free-floating in interstellar space (Sumi et al.
2011).
The solar system, with the widely spaced and nearly circular orbits of the giant planets,
bears little resemblance to the bulk of known exoplanets. Yet, if our understanding of physics
of planet–gas-disk interaction is correct, it seems likely that the young solar system followed
the evolutionary path outlined above. Due to the convergent planetary migration in times
before the gas disk’s dispersal, each planet should have become trapped in a resonance with
its neighbor. Jupiter and Saturn, for example, were most likely trapped in the 3:2 resonance
(Masset & Snellgrove 2001, Morbidelli & Crida 2007, Pierens & Nelson 2008), defined as
PSaturn/PJupiter = 1.5, where PJupiter and PSaturn are the orbital periods of Jupiter and Saturn
(this ratio is 2.49 today).
To stretch to the present, more relaxed state, the outer solar system most likely un-
derwent a violent phase when planets scattered off of each other and acquired eccentric
orbits (Thommes et al. 1999, Tsiganis et al. 2005). The system was subsequently stabilized
by damping the excess orbital energy into the transplanetary disk, whose remains survived
to this time in the Kuiper belt. Finally, as evidenced by dynamical structures observed
in the present Kuiper belt, planets radially migrated to their current orbits by scattering
planetesimals (Malhotra et al. 1995, Levison et al. 2008).
2. Method
We conducted computer simulations of the early evolution of the solar system in an
attempt to determine the initial states of planetary orbits (Batygin & Brown 2010). First,
we performed hydrodynamic and N-body simulations to identify the resonant configurations
that may have occurred among the young solar system’s giant planets. Our hydrodynamic
simulations used Fargo (Masset 2000) and followed the method described in Morbidelli et al.
(2007). As the Fargo simulations are CPU expensive, we used these results as a guide, and
generated many additional resonant systems with the N-body integrator known as SyMBA
(Duncan et al. 1998).
Planets with masses corresponding to those of Jupiter, Saturn and ice giants were placed
in initial orbits with the period ratios slightly larger that those of the selected resonances.
The planets were then migrated into resonances with the SyMBA code that was modified to
include forces that mimic the effects of gas. We considered cases with four and five initial
– 3 –
planets, where in the later case, an additional planet was placed into a resonant orbit between
Saturn and the ice giants, or beyond the orbit of the outer ice giant. The fifth planet was
given the mass between 3 × 1025 and 3 × 1026 g, which roughly corresponds to 1/3 and 3
times the mass of Uranus (or Neptune).
Different starting positions of planets, rates of the semimajor axis and eccentricity evo-
lution, and timescales for the gas disk’s dispersal produced different results. For Jupiter
and Saturn, we confined the scope of this study to the 3:2 and 2:1 resonances, because the
former one is strongly preferred from previous hydrodynamic studies (Masset & Snellgrove
2001, Morbidelli et al. 2007, Pierens & Nelson 2008). The 2:1 resonance was included for
comparison. The eccentricities and resonant amplitudes obtained here were typically e < 0.1
and < 60◦. The inner ice giant had 0.05 < e < 0.1 in most cases, while the other planets
acquired more circular orbits.
The instability of a planetary system can occur after the gas disk’s dispersal when the
stabilizing effects of gas are removed. Such an instability can be triggered spontaneously
(e.g., Weidenschilling & Marzari 1996), or by divergent migration of planets produced by their
interaction with planetesimals leaking into the planet-crossing orbits from the transplanetary
disk (Thommes et al. 1999). In the later case, the instability is produced when Jupiter and
Saturn cross a major mean motion resonance (Tsiganis et al. 2005; also known as the Nice
model).
In the solar system, it is often assumed that the instability occurred at the time of the
Late Heavy Bombardment (LHB) of the Moon some 3.9 Gy ago, when the lunar basins with
known ages formed (Hartmann et al. 2000). If so, the solar system’s giant planets would
be required to remain on their initial resonant orbits for about 700 Myr. To allow for this
possibility, we sifted through the resonant configurations identified above and selected those
that were stable over 109 yr, if considered in isolation. Only the stable systems were used
for the follow-up simulations, in which we tracked the evolution of planetary orbits through
and past the instability.
We included the effects of the transplanetary planetesimal disk in these simulations.
The disk was represented by 1000 equal-mass bodies that were placed into orbits with low
orbital eccentricities and inclinations, and in radial distances between rin < r < rout. The
surface density was set to be Σ = 1/r. The outer edge of the disk was placed at rout = 30
AU, so that the planetesimal-driven migration is expected to park Neptune near its present
semimajor axis.
We considered cases with rin = 0.5, 1, and 3.5 AU. The instability was triggered early
for rin = 0.5 AU, as planetesimals leaked from the inner part of the disk into the planetary
– 4 –
region, and interacted with planets. To trigger the instability for rin = 1 and 3.5 AU, we
broke the resonant locks by altering the mean anomaly of one of the ice giants. This method
was inspired by the recent study of the late instability that showed that planets can slowly
exchange orbital energy with a distant planetesimal disk, and break from the resonances
when the resonant amplitude exceeds certain limits (Levison et al. 2011).
In either case discussed above, the scattering phase between planets starts shortly after
the beginning of our simulations, which guarantees low CPU cost. We considered four
different masses of the planetesimal disk, mdisk = 10, 20, 35, 50, 75 and 100 MEarth, and
performed 30 simulations in each case, where different evolution histories were generated
by randomly seeding the initial orbit distribution of planetesimals. In total, we completed
over 6000 scattering simulations. Each system was followed for 100 Myr, at which point the
planetesimal disk was largely depleted and planetary migration ceased.
3. Constraints
We defined several criteria to measure the overall success of simulations. First of all, the
final planetary system must have four giant planets (criterion A) with orbits that resemble
the present ones (criterion B). Note that A means that one planet must be ejected in the
five-planet runs, while all four planets need to survive in the four-planet runs. As for B, we
claim success if the final mean semimajor axis of each planet is within 20% to its present
value, and if the final mean eccentricities and mean inclinations are no larger than 0.11 and
2◦, respectively. These thresholds were obtained by doubling the current mean eccentricity
of Saturn (eSaturn = 0.054) and mean inclination of Uranus (iUranus = 1.02◦).
For the successful runs, as defined above, we also checked on the history of encounters
between giant planets, evolution of the secular g5, g6 and s6 modes, and secular structure
of the final planetary systems. To explain the observed populations of the irregular moons
that are roughly similar at each planet (Jewitt & Haghighipour 2007), all planets –including
Jupiter– must participate in encounters (Nesvorn´y et al. 2007). Encounters of Jupiter and/or
Saturn with ice giants may also be needed to excite the g5 mode in the Jupiter’s orbit to its
current amplitude (e55 = 0.044; Morbidelli et al. 2009).
It turns out that it is generally easy to have encounters of one of the ice giants to Jupiter
if planets start in a compact resonant configuration. The amplitudes of g6 and s6 modes
also do not pose a problem.
It is much harder to excite e55, however. We therefore opt
for a criterion in which we require that e55 > 0.22 in the final systems, i.e., at least half
of its current value (criterion C). The e55 amplitude was determined by following the final
– 5 –
planetary systems for additional 10 Myr (without planetesimals), and Fourier-analyzing the
results (Sidlichovsk´y & Nesvorn´y 1996).
The evolution of secular modes, mainly g5, g6 and s6, is constrained from their effects
on the terrestrial planets and asteroid belt. As outer planets scatter and migrate, these
frequencies change. This may become a problem, if g5 slowly swipes over the g1 or g2 modes,
because the strong g1 = g5 and g2 = g5 resonances can produce excessive excitation and
instabilities in the terrestrial planet system (Brasser et al. 2009, Agnor & Lin 2011). The
behavior of the g6 and s6 modes, on the other hand, is important for the asteroid belt
(Morbidelli et al. 2010).
As g5, g6 and s6 are mainly a function of the orbital separation between Jupiter and
Saturn, the constraints from the terrestrial planets and asteroid belt can be conveniently
defined in terms of PSaturn/PJupiter. This ratio needs to evolve from <2.1 to >2.3 in < 1 Myr
(criterion D), which can be achieved, for example, if planetary encounters with an ice giant
scatter Jupiter inward and Saturn outward. This condition may therefore require planetary
encounters, but is more subtle in that not all simulations with Jupiter encounters are good.
4. Results
4.1. Four- and Five-Planet Results
We start by discussing the results obtained with four planets that were assumed to be
initially locked in the (3:2,3:2,4:3) resonances. The inner ice giant has the largest eccentricity
(e3 = 0.06). According to Levison et al.
(2011), the system should therefore be driven
toward the instability by the energy and momentum exchange between the inner ice giant
and planetesimal disk. We set rin = 15 AU, so that the inner disk edge is well beyond the
orbits of the two ice giants (a3 = 9.6AU and a4 = 11.6 AU). This setup should be consistent
with the late instability (Levison et al. 2011).
The best results were obtained with Mdisk = 35MEarth and Mdisk = 50MEarth. The
fraction of simulations producing the final systems with four outer planets is 10% and 13%,
respectively (criterion A). Only three of the total of 120 integrations performed for these
disk masses satisfied our criterion B. This shows that it is very unlikely that the solar system
evolved from these initial states. The results do not improve when Mdisk is varied. The light
disks with Mdisk ≤ 20MEarth lead to final systems with fewer than 4 planets. The heavy
disks with Mdisk > 50MEarth do not work as well, as we discuss in more detail below (§4.2).
We compare these results with those obtained for the five-planet systems. To start with,
– 6 –
the five planets were placed into the (3:2,3:2,4:3,5:4) resonances and fifth planet was given
the mass equal to that of Uranus. As in the four-planet case, we fix rin = 15 AU so that the
most eccentric inner ice giant (e3 = 0.07) has about the same radial distance from the inner
edge of the disk.
The best five-planet results were obtained for Mdisk = 50MEarth with the fraction of
systems matching criteria A and B raising to 37% and 23%, respectively (Fig. 1). This shows
that it is it roughly ten times more likely to obtain a good solar system analog starting from
five planets than from four planets, at least in the case discussed above. This result is not
unexpected because systems starting from a compact resonant configuration typically suffer
a rather violent instability, and tend to loose planets (Fig. 2).
4.2. Effects of Mdisk
The results for the four-planet case do not improve when the mass of the planetes-
imal disk is increased. With Mdisk = 100MEarth, Neptune migrates too far and/or the
divergent migration of Jupiter and Saturn moves these planets too far apart so that 2.8 <
PSaturn/PJupiter < 3.2 in the end. The former problem can be resolved if the disk were trun-
cated at ≃25 AU, but we do not see any obvious cure to the second problem. This is because
the final radial spacing of the Jupiter’s and Saturn’s orbits depends on the mass of material
processed through <10 AU, which is simply too large for Mdisk = 100MEarth. In addition,
Saturn and Jupiter tend to end up on unrealistically circular orbits for large Mdisk, because
their eccentricities are damped by the large mass that these planets interact with.
4.3. Effects of rin
The success rate of simulations depends on rin with the larger rin values leading to a lower
success rate. For example, the best four-planet case discussed above matched the criterion
B in 10% of cases for rin = 12.6 AU and Mdisk = 50MEarth, which is only slightly lower that
the success rate obtained for the five-planet case with rin = 17.5 AU and Mdisk = 50MEarth.
On the other hand, the success rate of the four-planet case drops to .3% if rin = 17.5 AU.
The sensitivity of the success rate to rin stems from the following. With the small
rin values, planetesimals can rapidly leak from the inner part of the transplanetary disk
into the planetary region, and stabilize planets very soon after the onset of the instability.
With large rin, on the other hand, the system lacks the stabilizing effect of planetesimals
on planet-crossing orbits, planetary orbits become increasingly excited, and planets can be
– 7 –
ejected. This highlights the importance of planet-crossing population of planetesimals on
the results.
4.4. Mass of the Fifth Planet
We tested cases with different masses of the fifth planet. Following the mass gradient,
we placed the fifth planet with the mass intermediate between those of Saturn and Uranus
into an exterior resonance with Saturn, or a planet with mass lower than that of Neptune
beyond the orbit of Neptune. The former case did not lead to any improvement of the
results discussed above, because the system became to violent and frequently ejected the
less massive planets from the system. In the later case, the statistics also did not improve
because the inner ice giant got ejected, thus leaving a final system with incorrect masses.
4.5.
Initial Resonant Configuration
We found that the behavior of planetary systems, and whether or not the final systems
end up resembling the outer solar system, do not depend in detail on the initial resonant
sequence between Saturn and ice giants.
Instead, the results are mainly sensitive to the
initial resonance between Jupiter and Saturn, with 3:2 or 2:1 cases considered here, and the
overall initial spread of the ice giants’ orbits.
As for the 3:2 Jupiter-Saturn resonance, the case with (3:2,3:2,4:3,5:4) discussed above
was one of the most compact resonant systems studied here. Other compact systems show
similar success rates. The most relaxed system studied here was (3:2,3:2,3:2,3:2)=(3:2)3.
With rin = 17.5 AU (a5 = 16.1 AU in this case) and Mdisk = 35MEarth, the success rate
was higher than the one obtained for a similar setup with (3:2,3:2,4:3,5:4). For example, the
criterion A was satisfied in 33% of simulations with (3:2)3, while it was 13% in the former
case, and criterion B was satisfied in 17% of simulations. This shows that a more relaxed
initial resonant configuration of ice giants can improve the statistics (mainly for the lower
mass disks; Mdisk . 35MEarth).
The four-planet case with the 2:1 Jupiter-Saturn resonance shows a relatively large
success rate in matching our criteria A and B. The instability produced in these simulations
tends to be weaker than in the 3:2 case, which seems to have positive impact on the results.
There are fewer cases of Jupiter encounters with one of the ice giants, however, leading to
orbit histories that violate constraints C and D. We did not found any case, for (2:1,3:2,3:2)
and Mdisk = 35MEarth, where these criteria would be satisfied.
– 8 –
The five-planet case with the 2:1 Jupiter-Saturn resonance is interesting. Better results
were obtained in this case when the inner ice giant had lower mass, because when Jupiter
and Saturn start in the 2:1 resonance, their period ratio needs to change by ∼0.5 only (from
2 to 2.49), which requires a smaller perturbation. Assuming a planet half the Uranus mass,
20 ≤ Mdisk ≤ 35MEarth and rin = a5 + 1 AU, where a5 denotes the semimajor axis of the
outer ice giant, criterion A was satisfied in ∼50% of cases and criterion B was satisfied in
20-30% of cases. These results are pretty much independent of the initial resonant sequence
between Saturn and the ice giants. All successful jobs show Jupiter encounters and have
∼ 10% chance of simultaneously matching criteria C and D as well (§4.8).
4.6. Planetary encounters
Only ∼3% of simulations performed for the four-planet case with (3:2,3:2,4:3), rin = 12.6
AU, and Mdisk = 50MEarth, satisfy criteria A and B, and show encounters of one of the
ice giants to Jupiter. The statistics for other resonances is similarly low. Consequently,
in absence of Jupiter encounters, Jupiter and Saturn end up too close to each other if
Mdisk < 50MEarth, and their orbits are too circular.
Larger disk masses produce more plausible final period ratios, as Jupiter and Saturn can
slowly separate from each other by scattering planetesimals, but these evolution paths do
not satisfy criterion D. Overall, the criterion D was satisfied only in ∼1% of the four-planet
cases, which is troubling. For a comparison, practically in all five-planet simulations reported
here, all planets, including Jupiter, participate in planetary encounters. For example, the
criterion D was satisfied in 50% of cases that also satisfied A and B for (3:2,3:2,4:3,5:4),
Mdisk = 50MEarth and rin = 15 AU (Fig. 3), and in over 60% of good cases for (2:1,3:2,3:2,4:3),
Mdisk = 20MEarth and rin = 19 AU.
4.7. Secular Modes
We analyzed our simulations to determine the fraction of cases in which the secular
structure of the final systems was similar to that of the present solar system. We did not
use the Jupiter’s encounters as a proxy for good cases (Batygin & Brown 2010), because
we found that most simulations with Jupiter’s encounters did not satisfy our criterion C
(e55 > 0.22). In fact, the overall success rate for the criterion C was disappointingly low,
both for the four- and five-planet systems. In most cases, the initial eccentricity of Jupiter,
or the one excited by planetary encounters, was damped by planetesimals passing through
– 9 –
< 10 AU, so that e55 . 0.01 in the end.
This problem could most easily be resolved for low initial masses of the planetesimal
disk, because in this case, the e55 amplitude –initial or produced by the planetary encounters–
will most likely survive. As planets can easily be ejected for low Mdisk, more than four initial
planets will probably be required. Still, to make things work for the promising five-planet
case, the success in matching the criteria A and B should be improved for low Mdisk, because
the systems studied here with Mdisk = 20MEarth were generally too violent. This could be
achieved, for example, by including some sort of dissipation in the planetesimal disk, possibly
produced by the collisions between planetesimals.
5. Conclusions
The formation of Uranus and Neptune is a long-standing problem in planetary sci-
ence, because their accretion at ∼20 and ∼30 AU would require implausibly long timescales
(Safronov 1969, Levison & Stewart 2001). The ice giants can form more easily at <15 AU,
where densities are higher and dynamical timescales are shorter (e.g., Robinson & Boden-
heimer 2010, Jakub´ık et al. 2011). Our five-planet resonant systems start with all ice giant
at <15 AU, if Jupiter and Saturn are in the 3:2 resonance. If Jupiter and Saturn start in
the 2:1 resonance, on the other hand, the whole planetary system is more spread, and the
outer ice giant is at ∼18-20 AU, where its formation can be problematic. Future work will
need to address these issues.
This work was supported by NLSI and NSF.
REFERENCES
Agnor, C. B., & Lin, D. N. 2011, to be submitted
Batygin, K., & Brown, M. E. 2010, ApJ, 716, 1323
Brasser, R., Morbidelli, A., Gomes, R., Tsiganis, K., & Levison, H. F. 2009, A&A, 507, 1053
Cuk, M. 2007, Bulletin of the American Astronomical Society, 38, 537
Duncan, M. J., Levison, H. F., & Lee, M. H. 1998, AJ, 116, 2067
Hartmann, W. K., Ryder, G., Dones, L., & Grinspoon, D. 2000, Origin of the Earth and
Moon, 493
– 10 –
Jakubik, M., Morbidelli, A., Neslusan, L., & Brasser, R. 2011, arXiv:1107.2235
Jewitt, D., & Haghighipour, N. 2007, ARA&A, 45, 261
Kley, W. 2000, MNRAS, 313, L47
Levison, H. F., & Stewart, G. R. 2001, Icarus, 153, 224
Levison, H. F., Morbidelli, A., Vanlaerhoven, C., Gomes, R., & Tsiganis, K. 2008, Icarus,
196, 258
Levison, H. F., Morbidelli, A., Tsiganis, K., Nesvorn´y, D., Gomes, R., 2011, submitted to
AJ
Malhotra, R. 1995, AJ, 110, 420
Masset, F. 2000, A&AS, 141, 165
Masset, F., & Snellgrove, M. 2001, MNRAS, 320, L55
Morbidelli, A., & Crida, A. 2007, Icarus, 191, 158
Morbidelli, A., Tsiganis, K., Crida, A., Levison, H. F., & Gomes, R. 2007, AJ, 134, 1790
Morbidelli, A., Brasser, R., Tsiganis, K., Gomes, R., & Levison, H. F. 2009, A&A, 507, 1041
Morbidelli, A., Brasser, R., Gomes, R., Levison, H. F., & Tsiganis, K. 2010, AJ, 140, 1391
Nesvorn´y, D., Vokrouhlick´y, D., & Morbidelli, A. 2007, AJ, 133, 1962
Pierens, A., & Nelson, R. P. 2008, A&A, 482, 333
Robinson, S. E., & Bodenheimer, P. 2010, Bulletin of the American Astronomical Society,
42, 958
Safronov, V. S. 1969, in Evolution of the Protoplanetary Cloud and Formation of the Earth
and Planets (Nauka, Moscow)
Sidlichovsk´y, M., & Nesvorn´y, D. 1996, Celestial Mechanics and Dynamical Astronomy, 65,
137
Sumi, T., et al. 2011, Nature, 473, 349
Thommes, E. W., Duncan, M. J., & Levison, H. F. 1999, Nature, 402, 635
Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005, Nature, 435, 459
– 11 –
Weidenschilling, S. J., & Marzari, F. 1996, Nature, 384, 619
This preprint was prepared with the AAS LATEX macros v5.0.
– 12 –
Fig. 1.— Final orbits obtained with five planets starting in the (3:2,3:2,4:3,5:4) resonances,
Mdisk = 50MEarth, and rin = 15 AU. Only the systems ending with four planets are plotted
(dots). The error bars show the average and rms of the orbital elements. The mean orbits of
Jupiter, Saturn, Uranus and Neptune are shown by red, green, turquoise and blue triangles.
– 13 –
Fig. 2.— The orbit histories of giant planets in one of the simulation with five initial planets.
In this case, the five planets were started in the (3:2,3:2,4:3,5:4) resonances, Mdisk = 50MEarth,
and rin = 15 AU. After series of encounters with Jupiter, the inner ice giant was ejected from
the solar system at 8.2 × 105 yr (purple path). The remaining planets were stabilized by the
planetesimal disk, and migrated to orbits that very closely match those of the outer planets
(dashed lines).
– 14 –
Fig. 3.— Period ratio, PSaturn/PJupiter, for a selected five-planet case with (3:2,3:2,4:3,5:4)
and Mdisk = 35MEarth. The fifth planet was ejected by Jupiter at 3.5 Myr. This produced
a jump of PSaturn/PJupiter from ∼ 1.5 to 2.4. The kind of evolution shown here, known as
the jumping Jupiter (Morbidelli et al. 2010), may be needed to avoid the region at 2.1-2.3,
where the g1 = g5 and g2 = g5 resonances occur (Brasser et al. 2009, Agnor & Lin 2011).
|
1210.4579 | 1 | 1210 | 2012-10-16T21:29:01 | Dead Zones In Circumplanetary Discs as Formation Sites For Regular Satellites | [
"astro-ph.EP"
] | Regular satellites in the solar system are thought to form within circumplanetary discs. We consider a model of a layered circumplanetary disc that consists of a nonturbulent midplane layer and and strongly turbulent disc surface layers. The dead zone provides a favorable site for satellite formation. It is a quiescent environment that permits the growth of solid bodies. Viscous torques within the disc cause it to expand to a substantial fraction of its Hill radius (~0.4 R_H) where tidal torques from the central star remove its angular momentum. For certain parameters, the dead zone develops into a high density substructure well inside the Hill sphere. The radial extent of the dead zone may explain the compactness of the regular satellites orbits for Jupiter and Saturn. The disc temperatures can be low enough to be consistent with the high ice fractions of Ganymede and Callisto. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 7 (2011)
Printed 16 October 2018
(MN LATEX style file v2.2)
Dead Zones In Circumplanetary Discs as Formation Sites
For Regular Satellites
Stephen H. Lubow and Rebecca G. Martin
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
ABSTRACT
Regular satellites in the solar system are thought to form within circumplanetary discs.
We consider a model of a layered circumplanetary disc that consists of a nonturbulent
midplane layer and and strongly turbulent disc surface layers. The dead zone provides
a favorable site for satellite formation. It is a quiescent environment that permits
the growth of solid bodies. Viscous torques within the disc cause it to expand to a
substantial fraction of its Hill radius (∼ 0.4RH) where tidal torques from the central
star remove its angular momentum. For certain parameters, the dead zone develops
into a high density substructure well inside the Hill sphere. The radial extent of the
dead zone may explain the compactness of the regular satellites orbits for Jupiter and
Saturn. The disc temperatures can be low enough to be consistent with the high ice
fractions of Ganymede and Callisto.
Key words: accretion, accretion discs -- planets and satellites: formation -- planetary
systems
1
INTRODUCTION
reaches
a value
forms
circumplanetary disc
Lubow, Seibert & Artymowicz
After a planet forms in a circumstellar disc and be-
its mass
fore
of Jupiter's mass,
tidal
forces from the planet open a gap in the disc
(Lin & Papaloizou 1984, 1986; D'Angelo, Henning & Kley
2002; Bate et al. 2003). The planet continues to accrete
material from the circumstellar disc through the gap and
(Artymowicz & Lubow
a
1996;
1999;
D'Angelo, Henning & Kley 2003; Ayliffe & Bate 2009).
A circumplanetary (spin-out) disc may also form if the
planet manages to achieve break-up velocities as it contracts
(Korycansky, Bodenheimer & Pollack 1991). The regular
satellites in the solar system have prograde, nearly circular,
and nearly coplanar orbits. They are thought to have formed
in a circumplanery disc, in analogy with the formation of
planets in the solar nebula (Pollack, Lunine & Tittemore
1991). Understanding the structure of these discs is essential
for explaining regular satellite formation.
There are several key constraints that various models
of gaseous circumplanetary discs have attempted to satisfy.
The regular satellites around Jupiter and Saturn lie within a
radius of less than 0.06RH of their Hill (tidal) radii RH. The
small extent of this region has been explained to be a conse-
quence of a compact disc structure or substructure. Various
studies have tied the compactness to the level of angular mo-
mentum of accreting material entering a planet's Hill sphere
(e.g., Estrada et al. 2009; Ward & Canup 2010). The idea is
c(cid:13) 2011 RAS
that if the accreting inflow has sufficiently low angular mo-
mentum per unit mass, then the satellites formed with a
circumplanetary disc will occupy a small region within the
Hill sphere.
All hydrodynamics simulations of gas flows about a
Jupiter mass planet have revealed a disc that is consider-
ably more extended than the Gallilean satellites. The recent
study by Ayliffe & Bate (2009) shows that the disc extends
to about 0.35RH. For a planet such as Jupiter that opens
a gap in the circumstellar disc, gas enters the Hill sphere
preferentially near the Lagrange points and carries signifi-
cant angular momentum. However, the main issue is that
the compactness of a fully turbulent disc cannot be main-
tained due to the action of viscous torques (Ward & Canup
2010; Martin & Lubow 2011a). These torques cause a com-
pact circum-Jovian disc to rapidly expand in only ∼ 100
planet orbital periods for typical disc parameters.
Although the angular momentum of the inflowing gas
is uncertain, the radial extent of the disc is largely inde-
pendent of this quantity (Martin & Lubow 2011a). The disc
expands to a radius where the tidal torques from the cen-
tral star (Sun) remove angular momentum at the rate it
is supplied by inflowing gas. This radius is estimated as
about 0.4RH, where ballistic periodic orbits begin to cross
(Martin & Lubow 2011a). (Some departures from this value
occur due to pressure effects.) In addition, no compact sub-
structure is found to lie within such a circumplanetary disc
(Estrada et al. 2009; Ayliffe & Bate 2009; Martin & Lubow
2011a). With this simple alpha disc model, the turbulent
2
S. H. Lubow & R. G. Martin
viscosity is by assumption a smooth function of radius. The
disc structure then follows that of a standard smooth accre-
tion disc.
Previous work such as Canup & Ward (2002) suggested
that if the inflowing angular momentum is small enough, the
flow may penetrate deep into the Hill sphere before initially
collecting into a disc. If the disk is viscous, its gas compo-
nent will then spread outward to a large radius. But the
solids initially delivered with the gas may not be effectively
coupled to the gas if they rapidly accrete and grow once in
circumplanetary orbit. The solids then accrete near the com-
pact region where they were initially delivered, rather than
tracking the outward viscous expansion of the gas disk.
There are some issues of concern with this picture. The
inflow may be largely planar and join the circumplanetary
disk at its outer edge, as indicated by simulations by Ayliffe
& Bate (2009). The gas and coupled solids would then not
be able to penetrate deeply within the Hill sphere. Instead,
they achieve Keplerian orbits near the outer edge of the pre-
existing disk as they becomes entrained there. It is possible
that larger size solid bodies that are decoupled from the gas
could enter the Hill sphere with sufficiently low angular mo-
mentum to be captured close to the planet. But it is not clear
that such a low angular momentum occurs. In this paper,
we suggest an alternative explanation for the compactness
of the satellites based on the radial extent of a dead zone, a
region of low disc turbulence.
Another constraint applies to the disc temperature.
This snow line plays an important role in the composition of
forming satellites. Outside of the snow line, the solid mass
density is much higher because of water ice condensation.
The snow line occurs at a temperature, Tsnow, that is around
170 K (Hayashi 1981; Lecar et al. 2006). Ganymede and Cal-
listo, the two outer Gallilean satellites contain a substantial
fraction of ice (∼ 50%), while the two inner satellites contain
much less. The snow line of the disc is then required to have
been near the orbits of these outer satellites. The partially
differentiated structure of Callisto suggests that its ice never
fully melted and that the snow line was always inside its or-
bit (Lunine & Stevenson 1982). Since the disc temperature
increases with mass accretion rate, this constraint limits that
rate. This requirement implies that the disc accretion rates
be sufficient low, substantially lower than what would be ex-
pected during the T Tauri accretion phase (Canup & Ward
2002; Estrada et al. 2009). The satellite formation epoch is
therefore expected to occur late in the life of the solar neb-
ula, when it has lost a substantial fraction of its mass.
Another issue involved in the formation of both planets
and satellites in gaseous discs is the permitted level of tur-
bulence. Estimates of disc turbulence from simulations and
properties of observed systems suggest that relatively high
levels of turbulence α ∼ 0.01 − 0.3 are expected in fully
turbulent discs (e.g., King, Pringle & Livio 2007). Larger
solid bodies may form from the gravitational instability of a
dense layer of dust that settles near the disc mid-plane (e.g.,
Goldreich & Ward 1973). Even modest levels of turbulence,
involving α ≪ 0.01, can prevent solids from setting to a thin
enough layer near disc mid-plane for gravitational instability
to operate (Dubrulle et al. 1995; Cuzzi & Weidenschilling
2006). An alternative model for the growth of solids involves
the concentration of dust in turbulent eddies (Cuzzi et al.
2008). This model also adopted a weaker level of turbulence.
Another effect is that turbulence can cause destructive colli-
sions among larger solid bodies, thereby preventing growth
to the desired size (e.g., Ida, Guillot, & Morbidelli 2008).
These results suggest that lower values of disc turbulence
may be more favorable for satellite formation than is occurs
in a fully turbulent disc.
In a layered disc, the magnetic turbulence (MRI;
Balbus & Hawley 1991) is not active at all heights (Gammie
1996). Magnetic turbulence requires a certain level of ioniza-
tion for the gas to be well enough coupled to the magnetic
field. In sufficiently cool disc regions, temperatures are too
low to provide the needed level of the ionization for magnetic
turbulence to operate. Instead, the ionization is provided by
external sources of radiation, such as X-rays or cosmic rays.
If the disc surface density is high enough, this radiation may
not penetrate deep enough below the disc surfaces to pro-
vide turbulence at all heights. A so-called dead zone with
little or no turbulence develops around the disc midplane.
Such an environment may be favorable for the survival and
growth of solid bodies.
In an earlier paper, we explored a layered disc model for
the evolution of a circumplanetary disc during a stage when
the disc experiences accretion rates comparable to those of
the T Tauri phase (Lubow & Martin 2012). Such discs can
sometimes undergo outbursts, analogous to the FU Ori out-
bursts. In any case, such a disc is too hot for the survival of
icy satellites at the locations of Ganymede and Callisto.
We explore in this paper the viability of a layered disc
model as an environment for satellite formation in later
stages of its evolution. We do not attempt to model the
formation of satellites in such a disc. Instead, we show that
in this stage and for certain plausible disc ionization pa-
rameters, a layered disc model satisfies the requirements de-
scribed above. In particular, a compact high density disc
substructure develops in a low turbulence dead zone whose
extent matches that of the regular satellites. The disc tem-
perature is low enough for the outermost Galilean satellites
to lie outside the snow line.
In Section 2, we describe the results from our circum-
planetary disc model and how it may explain some features
of the formation of the Galilean satellites. In Section 3 we
discuss some implications and limitations of the model and
in Section 4 we draw our conclusions.
2 CIRCUMPLANETARY DISC MODEL
In this Section we first describe our layered circumplanetary
disc model that is truncated by tides from the star and then
we discuss the results.
2.1 Model Description
the
layered disc model
We use
initially described
in Armitage, Livio, & Pringle (2001) and further devel-
oped in Zhu et al. (2009), Martin & Lubow (2011b) and
Lubow & Martin (2012). Material in the circumplanetary
disc orbits the central planet of mass Mp at Keplerian an-
gular speed Ω(R) = pGMp/R3 for radius R from the
Minfall. The
planet. Material is added to the disc at rate
disc has a total surface density Σ(R, t), midplane temper-
ature Tc(R, t), and surface temperature Te(R, t) at time t.
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
The MRI turbulent surface layer (which we call the active
layer) has surface density Σm(R, t), temperature Tm(R, t),
and turbulent viscosity νm(Tm, R, t) that is parametrised
with the Shakura & Sunyaev (1973) αm parameter. Cosmic
rays and/or X-rays are assumed to be able to provide suf-
ficient ionization for MRI to operate to a constant surface
density depth Σcrit. If the total surface density is greater
than this critical value, Σ > Σcrit, then Σm = Σcrit. For
smaller surface density, Σ < Σcrit, the disc is fully MRI ac-
tive and Σm = Σ.
A complementary midplane layer exists if the total sur-
face density is larger than the critical, Σ > Σcrit. It has
surface density Σg = Σ − Σcrit and midplane temperature
Tc. This layer is MRI active if the midplane temperature is
greater than the critical, Tc > Tcrit, where Tcrit is the tem-
perature for sufficient thermal ionisation. However, for lower
temperatures, Tc < Tcrit, a dead zone forms. In the model,
there is no turbulence in the dead zone unless it is massive
enough to be self-gravitating. (In reality, some turbulence is
expected due to the vertical propagation of waves from the
active layer (Fleming & Stone 2003).) The condition for self
gravity is taken to be that the Toomre parameter is smaller
than its critical value, Q < Qcrit = 2 (Toomre 1964). In this
case, a second viscosity term is included in the complemen-
tary layer.
We consider the planet Jupiter with mass, Mp = 1 MJ,
that is orbiting the Sun, M = 1 M⊙, at a distance of a =
5.2 AU. We solve the accretion disc equations numerically on
a fixed mesh that is uniform in log R with 120 grid points
(e.g., Armitage, Livio, & Pringle 2001; Martin et al. 2007).
The inner boundary has a zero torque condition at the radius
of Jupiter, R = 1 RJ, so the mass falls freely on to the planet.
At the outer boundary we approximate the tidal torque from
the star with a zero radial velocity boundary condition at
Rout = 0.4 RH. Material accretes on to the disc at a constant
Minfall at a radius which we take to be 0.33 RH (see
rate
Lubow & Martin 2012, for more explanation). Initially, we
take the surface density and temperature structure of the
disc to be that of low mass (nonself gravitating) disc with a
small dead zone. This state represents the disc immediately
following an outburst. Alternatively, if there is initially no
mass in the disc, it quickly builds up to the same state.
The critical surface density that is sufficiently ionised to
be turbulent, Σcrit, is not well determined. When cosmic rays
are the dominant source of ionisation, for typical parameters
the surface turbulent layer of a circumstellar disc has been
estimated to have a surface density of Σcrit ≈ 200 g cm−2
(Gammie 1996; Fromang, Terquem & Balbus 2002). How-
ever, it is possible that cosmic rays are swept away from the
disc by means of a turbulent MHD jet or outflow which
may be present at the star (e.g. Skilling & Strong 1976;
Cesarsky & Volk 1978). In the absence of cosmic rays, X-
rays from the star may be the dominant source of ioni-
sation and in this case the active layer is much smaller
(Matsumura & Pudritz 2003).
More recent works find the inclusion of polycyclic aro-
matic hydrocarbons and dust tends to suppress the instabil-
ity further (Bai & Goodman 2009; Perez-Becker & Chiang
2011; Bai 2011; Fujii, Okuzumi & Inutsuka 2011). However,
in circumstellar discs, such effects produce an accretion
rate that is too low to account for T Tauri accretion rates
and this problem remains unsolved. Perez-Becker & Chiang
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
Dead Zones as Sites for Satellite Formation
3
(2011) suggest that in order to explain the T Tauri rates,
Σcrit > 10 g cm−2, in the circumstellar disc. It is possible
that circumplanetary discs may have an even smaller ac-
tive layer surface density than circumstellar discs, since they
may be shielded from the ionising radiation of the central
star by the shadowing effects of the inner circumstellar disc.
Because of these uncertainties, we follow the work of Ar-
mitage et al (2001) and Zhu et al. (2009) and regard Σcrit
as a constant free parameter. If the critical surface den-
sity is large, then the entire disc will be fully turbulent.
If it is small, then the dead zone will be very extensive.
We consider a range of active layer surface densities with
1 g cm−2 6 Σcrit 6 100 g cm−2.
Generally, we take the value of the temperature for
sufficient thermal
ionization for MRI to operate to be
Tcrit = 800 K. At this temperature the ionisation fraction
increases exponentially with temperature due to the colli-
sional ionisation of potassium (Umebayashi 1983). The vis-
cosity αm parameter associated with MRI is found to be
& 0.01 from MHD simulations, but depends on numerous
parameters such as the resolution, stratification, and treat-
ments of small scale dissipation and radiation transport
(e.g. Hartmann et al. 1998; Fromang et al. 2007; Guan et al.
2009; Davis et al. 2009). Similarly, observations of FU Ori
suggest that αm ≈ 0.01 (Zhu et al. 2007). However, obser-
vations of X-ray binaries and dwarf novae suggest αm ≈
0.1−0.4 (King, Pringle & Livio 2007). Armitage et al (2001)
take αm = 0.01 and Zhu et al. (2010) consider both 0.01 and
0.1. With the uncertainty in the value, we also consider cases
with both αm = 0.01 and 0.1.
Minfall,
During times of planet formation, the accretion rate
on to the circumplanetary disc,
is of order
the overall circumstellar disc accretion rate (Bate et al.
2003; Lubow & D'Angelo 2006; Ayliffe & Bate 2009) that
is of order 10−8 M⊙ yr−1 (Valenti, Basri & Johns 1993;
Hartmann et al. 1998), typical for the T Tauri phase. During
this phase, disc temperatures are too high to permit the sur-
vival of ice in Callisto (Canup & Ward 2002; Estrada et al.
2009). Consequently, we consider the evolution in the later
stages of disc evolution, when the accretion rate has dropped
by an order of magnitude or more. We consider accretion
rates in the range 10−11
− 10−9 M⊙ yr−1 near the end of the
disc lifetime.
If the circumplanetary disc density in the dead zone
grows sufficient large during the T Tauri stage, the disc be-
comes gravitationally unstable. Self-gravity generates tur-
bulence that raises the disc midplane temperature (e.g.,
Lodato & Rice 2004). If the temperature exceeds the crit-
ical value Tcrit for sufficient thermal
ionization for MRI
to operate, MRI can set in abruptly and lead to a
gravo-magneto accretion outburst in a circumplanetary disc
(Lubow & Martin 2012). These conditions involve high disc
temperatures. The dead zone structure is disrupted and
dead zone mass is accreted on the central planet. The pro-
cess repeats and can be described as a limit cycle of the
gravo-magneto instability (Martin & Lubow 2011b). Con-
sequently, we regard the T Tauri stage as unfavorable for
satellite formation or survival.
As the disc disperses near the end of its lifetime, the
Minfall decreases and outbursts become less
accretion rate
likely. We are most interested in models that have a long
outburst interval timescale. The outburst interval timescale
4
S. H. Lubow & R. G. Martin
increases with decreasing accretion rate. If the mass accre-
tion rate decreases substantially after an outburst, then the
outburst may not recur. The lifetime of the circumplanetary
disc is assumed to be similar to that of the circumstellar disc,
a few 106 yr. A simple estimate for the outburst timescale
is the timescale for the disc to become self gravitating. This
timescale can be estimated as the time to build up a mass of
MpH/R, where H = cs/Ω is the disc scale height and cs is
the sound speed. This mass is at most around a few tenths
of a Jupiter mass. The outburst interval timescale is of order
tint ≈
Mp
Minfall
H
R
.
(1)
2.2 Model Results
Table 1 summarizes simulation results for a range of pa-
rameters that cover some plausible values for the viscos-
ity parameter αm and the critical surface density for ion-
ization Σcrit. In Fig. 1 we show the surface density and
temperature distributions for models with αm = 0.01 and
Minfall = 10−10 M⊙ yr−1. Model R8 is fully turbulent and
this steady solution holds for Σcrit > 27.1 g cm−2. Model R9
has critical surface density Σcrit = 10 g cm−2 and hence con-
tains a dead zone. Within the dead zone, the surface density
increases substantially with time. For Model R9 this region
extends to radius Rdead = 0.034 RH. This location is close to
the orbit of Callisto. Such a model may explain why Callisto
has a large mass and angular momentum, yet no satellites
are found outside of its orbit. Beyond the dead zone, less
material is available to form satellites and the turbulence
may disrupt the growth into larger bodies.
At later times, the higher disc mass in the dead zone
is in the range of the so-called minimum mass sub-nebula
(MMSN) (Lunine & Stevenson 1982). However, the model
here is fundamentally different from a static MMSN that
was initially envisioned by some earlier work, because it con-
siders a continuously supplied disc. The accretion time scale
in this paper is regulated by the inflow rate to the disc, as
in the original continuous inflow concept of (Canup & Ward
2002).
The disc structure we obtain is similar to the gas rich
disc model envisioned by Mosqueira & Estrada (2003a) and
Estrada et al. (2009). In their model, Callisto lies just out-
side the dense inner disc, the dead zone in the current
model. They suggested that Callisto's inferred long for-
mation timescale (Stevenson, Harris & Lunine 1986), which
caused its partial differentiation, results from the rela-
tively slow delivery of solids in the outer disc, due to
its lower density. The model also describes Ganymede as
being formed more rapidly within the dense disc. These
different environments provide a possible explanation for
the Callisto-Ganymede dichotomy that describes the dif-
ferences between these two outermost Gallilean satellites
(Lunine & Stevenson 1982).
The mass of the fully turbulent disc in Model R8 is
4.3 × 10−5 MJ whereas the mass of the disc with a dead
zone in Model R9 at a time of 105 yr is 9.1 × 10−3 MJ. The
dead zone model has a lower temperature distribution than
the fully turbulent model. For example, the snow line loca-
tion changes from 0.032 RH in the fully turbulent model to
0.017 RH when a dead zone is introduced. The latter snow
line location is in the range expected for the survival of ice
in Callisto. The outburst intervals for Model R9 at constant
accretion rate are also fairly long. For model R9, H/R < 0.2
and so the outburst timescale is around 2 × 106 yr.
With a higher accretion rate of 10−9 M⊙ yr−1, the disc
builds up mass very quickly and so the outburst time scale
is fairly short. In addition, it is difficult to simultaneously
meet the constraints on the snow line location and the com-
pactness of the dead zone. The models that best meet these
requirements, Models R9 and R12, have the intermediate
accretion rate M = 10−10 M⊙ yr−1.
We note that because of the dead zone, the vertically
averaged α in the disc is much lower than the value in the
active layer of 0.01 or 0.1. In some ways this vertical aver-
age approximates the α value adopted by Canup & Ward
(2002). However, since that model does not have a dead
zone, there is no accumulation of mass that produces the
compact disc substructure. Fully turbulent models with
lower accretion rates of 10−11 M⊙ yr−1 (R13 and R14) have
low enough temperatures for icy satellite formation (as
found by Canup & Ward 2002).
3 DISCUSSION
There are several approximations in the layered disc model
we have considered that should be improved on in future
work. Some of these were discussed in Martin & Lubow
(2011b) and Lubow & Martin (2012). For example, we
have approximated the surface density in the active layer
as a constant with radius in the disc. An alternative
method of finding the dead zone is with a critical mag-
netic Reynolds number (e.g. Fromang, Terquem & Balbus
2002; Matsumura & Pudritz 2003). In this approach, the
surface density of the turbulent layer may vary in radius
(Martin et al. 2012a). However, there are currently still
problems with this model, as it is difficult to explain T
Tauri accretion rates with a low critical surface density
(Martin et al. 2012b).
The infall accretion rates we considered are constant
in time. However, they are likely to decrease as the flow
through the circumstellar disc decreases. As the accretion
rate decreases, the active surface layer occupies an increasing
fraction of the vertical extent of the disc. When the accretion
rate drops to values of . 10−11M⊙ yr−1, the disc will become
fully turbulent with the adopted parameters (Models R13
and R14). The disc temperatures remain low enough for the
survival of ice in Callisto.
The inferred density of the small inner Jovian satel-
lite Amalthea suggests that contains a substantial fraction
of water-ice (Anderson et al. 2005). Accretion disc models,
such as the one described here, predict that temperatures
are too high for the water-ice to survive and requires some
other explanation.
Another open issue is satellite survival in such a high
density disc against the effects of orbit decay due to disc-
satellite interactions (Ward 1997; Lubow & Ida 2010). The
Type I migration timescale in the dead zone of model R9
is quite short ∼ 102yr (Canup & Ward 2002). It is possi-
ble that the satellites instead undergo the potentially slower
Type II migration, provided that they can open gaps. Gap
opening in a low viscosity disc is dominated by effects of gas
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
Dead Zones as Sites for Satellite Formation
5
Model
Minfall
(M⊙ yr−1)
R1
R2
R3
R4
R5
R6
R7
R8
R9
R10
R11
R12
R13
R14
10−9
10−9
10−9
10−9
10−9
10−9
10−9
10−10
10−10
10−10
10−10
10−10
10−11
10−11
αm
0.01
0.01
0.01
0.01
0.1
0.1
0.1
0.01
0.01
0.01
0.1
0.1
0.01
0.1
Σcrit
(g cm−2)
Rdead
(RH)
Rsnow Mdisc(t = 105 yr) Outburst Timescale
(RH)
(MJ)
(yr)
> 90.3
40
10
1
> 15.6
10
1
> 27.1
10
1
> 4.7
1
> 8.1
> 1.2
-
0.023
0.32
0.33
-
0.0097
0.32
-
0.034
0.32
-
0.035
-
-
0.094
0.084
0.019
0.0023
0.056
0.052
0.011
0.032
0.017
0.0023
0.021
0.011
0.011
0.0092
1.9 × 10−4
6.7 × 10−2
1.0 × 10−1
1.1 × 10−1
2.5 × 10−4
1.0 × 10−1
4.3 × 10−5
9.1 × 10−3
1.1 × 10−2
4.5 × 10−6
9.8 × 10−3
7.7 × 10−6
7.8 × 10−7
Steady
1.4 × 105
7.8 × 105
8.3 × 105
Steady
> 105
> 105
Steady
> 106
> 106
Steady
> 106
Steady
Steady
Table 1. Column 2 contains the infall accretion rate on to the circumplanetary disc, column 3 contains the viscosity αm parameter,
column 4 contains the critical surface density in the active layer that is ionised by cosmic rays/X-rays, column 5 contains the outer radius
of the dead zone, Rdead, if it exists, column 6 contains the snow line radius and column 7 contains the total mass in the disc at time
t = 105 yr. If the disc has a dead zone, then the mass of the disc increases linearly in time. But if there is no dead zone, then the mass
in constant. Finally, column 8 contains estimates of the timescales for the gravo-magneto outbursts.
Figure 1. Left: The surface density of the disc in model R8 (dashed line) and R9 (solid lines) at times t = 102, 103, 104, 105 and
106 yr in order of increasing height in the plot. As the dead zone gains mass, its surface density grows in time. Right: The midplane
temperature of the disc in R9 (solid line) and R8 (dashed line). The temperature in model R9 does not change in time even though the
surface density does. The dotted line shows the snow line temperature, Tsnow = 170 K. The solid circles in both plots show the radial
locations of the Galilean satellites.
shocks caused by the steepening of density waves (Rafikov
2002; Yu et al. 2010). The gap opening condition in this
case suggests that Callisto and Ganymede can open gaps
for Model R9, but only for Σ . 104g cm−2.
The nature of gap opening and migration in a layered
disc has not been examined. The application of the viscous
gap opening criterion (e.g., equation 5 of Canup & Ward
2002) to the active layer alone suggests that gap opening
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
does not occur. On the other hand, its application to a ver-
tical averaged α suggests that gap opening may occur. There
are some other possibilities. If Callisto resides just out-
side the dead zone, it may be subject to outward Lindblad
torques that stall its migration there (Mosqueira & Estrada
2003b; Matsumura, Pudritz & Thommes 2007). Trapping
may also be possible at the inner edge of the dead zone
(Kretke et al. 2009). We also note that outward corotation
6
S. H. Lubow & R. G. Martin
torques could affect migration (Paardekooper & Mellema
2006), since the dead zone region in Fig. 1 has a negative
radial entropy gradient that is required for this effect to op-
erate. On the other hand, coorotation torques can become
quite weak and saturate in a low viscosity environment (e.g.,
Ward 1991; Ogilvie & Lubow 2003).
Canup & Ward (2002) point out other constraints that
should be considered. One is that if satellites open gaps in
the disc, then satellite eccentricities should be excited by
first order Lindblad resonances. The excitation of eccentric-
ity of planets that open gaps has been an active area of ex-
ploration (Goldreich & Sari 2003; Ogilvie & Lubow 2003).
The simulation results appear to be sensitive to how clean
and large the gap is. Papaloizou. Nelson, & Masset (2001)
found no eccentriciy growth for a 1MJ which produces
a fairly clean gap, but growth at much higher masses.
D'Angelo, Lubow & Bate (2006) found some eccentricity
growth occurred 1MJ, but was quite weak compared to
growth rates for higher mass planets. In any case, at late
times when Minfall . 10−11M⊙ yr−1, the disc surface den-
sity in the satellite region will drop to less than critical values
Σ(R) < Σcrit. The disc is then fully turbulent and gap open-
ing will not occur. The available gas at this stage may be
sufficient to damp eccentricities developed in a prior open-
gap stage.
The disc
structure produced by this model
is
somewhat similar to the gas rich model described in
Mosqueira & Estrada (2003a) and Estrada et al. (2009). In
that model, the dense inner disc is the relic of a disc formed
prior to gap opening. The low density outer disc is formed
by higher angular momentum gas accreted after gap open-
ing. It is not clear why the inner disc would not accrete onto
the planet. We expect that a significant fraction of the mass
of Jupiter was acquired after gap opening through disc ac-
cretion. In the present model, accreting gas builds up mass
in a small inner region due to the presence of the dead zone.
The dead zone is assumed to be free of turbulence.
Suppose some independent mechanism produces turbulence
there that we do not take into account, such as a hydro-
dynamic instability not involving MRI or self-gravity. The
dead zone would then achieve a steady state flow that would
limit the mass growth of the dead zone as seen in Fig. 1. For
the parameters adopted in Model R9, reaching the highest
plotted densities requires that αd ∼ 10−8.
4 CONCLUSIONS
We have described some possible proto-satellite environ-
ments that occur in discs with dead zones. A dead zone
will occur if external radiation is insufficient to fully ionize
cool regions of a disc (Gammie 1996). The dead zone pro-
vides a quiescent environment for the survival and growth
of solid bodies into satellites. As the dead zone gains mass,
it can provide a high density, compact substructure within
the circumplanetary disc (see Fig. 1). The regular satellites
of Jupiter and Saturn lie within a small fraction of their Hill
radii. A fully turbulent simple alpha disc extends to much
larger radii and has a smoothly varying density structure.
The compactness of a high density dead zone provides a pos-
sible explanation for the small radial extent of the regular
satellites. For accretion rates appropriate to the late stages
its evolution, the circum-Jovian disc can be cool enough to
permit the survival of the ice in Callisto.
In this paper we have only considered possible disc
structures that can arise in circumplanetary discs with dead
zones. More work is required to explore the growth, survival,
and evolution of satellites in such discs.
ACKNOWLEDGEMENTS
acknowledges
support
SHL
grant
NNX11AK61G. RGM thanks the Space Telescope Sci-
ence Institute for a Giacconi Fellowship. We thank the
referee for helpful comments.
from NASA
REFERENCES
Anderson, J. D., Johnson, T. V., Schubert, G., et al. 2005, Sci-
ence, 308, 1291
Armitage P. J., Livio M., Pringle J. E., 2001, MNRAS, 324, 705
Artymowicz P., Lubow S. H., 1996, ApJ, 467, L77
Ayliffe B. A., Bate M. R., 2009, MNRAS, 393, 49
Balbus S. A., Hawley J. F., 1991, ApJ, 376, 214
Bai X. N., 2011, ApJ, 739, 50
Bai X. N., Goodman J. 2009, ApJ, 701, 737
Bate M. R., Lubow S. H., Ogilvie G. I., Miller K. A., 2003,
MNRAS, 341, 213
Canup R. M., Ward W. R., 2002, ApJ, 124, 3404
Cesarsky C. J., Volk H. J., 1978, A&A, 70, 367
Cuzzi, J. N., & Weidenschilling, S. J. 2006, Meteorites and the
Early Solar System II, 353
Cuzzi, J. N., Hogan, R. C., & Shariff, K. 2008, ApJ, 687, 1432
D'Angelo G., Henning T., Kley W., 2002, A&A, 385, 647
D'Angelo G., Henning T., Kley W., 2003, ApJ, 599, 548
D'Angelo, G., Lubow, S. H., & Bate, M. R. 2006, ApJ, 652, 1698
Davis S. W., Blaes O. M., Hirose S., Krolik J. H., 2009, ApJ,
703, 569
Dubrulle, B., Morfill, G., & Sterzik, M. 1995, Icarus, 114, 237
Estrada P. R., Mosqueira I. L., Lissauer J. J., D'Angelo G.,
Cruikshank D. P., 2009, in Europa, eds Pappalardo R. T.,
McKinnon W. B., Khurana K. K., University of Arizona Press,
Tucson, 27
Fleming, T., & Stone, J. M. 2003, ApJ, 585, 908
Fromang S., Terquem C., Balbus S. A., 2002, MNRAS, 329, 18
Fromang S., Papaloizou J., Lesur G., Heinemann T., 2007, A&A,
476, 1123
Fujii Y. I., Okuzumi S., Inutsuka S., 2011, ApJ, 2011, 743, 53
Gammie C. F., 1996, ApJ, 457, 355
Goldreich, P., & Ward, W. R. 1973, ApJ, 183, 1051
Goldreich, P., & Sari, R. 2003, ApJ, 585, 1024
Guan X., Gammie C. F., Simon J. B., Johnson B. M., 2009,
ApJ, 694, 1010
Hartmann L., Calvet N., Gullbring E., D'Alessio P., 1998, ApJ,
495, 385
Hayashi C., 1981, Prog. Theor. Phys. Suppl., 70, 35
Ida, S., Guillot, T. , Morbidelli, A., 2008, ApJ, 686, 1292.
King A. R., Pringle J. E., Livio M., 2007, MNRAS, 376, 1740
Korycansky, Bodenheimer, & Pollack 1991, Icarus, 92, 234
Kretke, K. A., Lin, D. N. C., Garaud, P., & Turner, N. J. 2009,
ApJ, 690, 407
Lecar M., Podolak M., Sasselov D., Chiang E., 2006, ApJ, 640,
1115
Lin D. N. C., Papaloizou J.,1986, ApJ, 309, 846
Lin D. N. C., Papaloizou J.,1984, ApJ, 285, 818
Lodato G., Rice W. K. M., 2004, MNRAS, 351, 630
Lubow S. H., Seibert M., Artymowicz P., 1999, ApJ, 526, 1001
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
Dead Zones as Sites for Satellite Formation
7
Lubow S. H., D'Angelo G., 2006, ApJ, 641, 526
Lubow S. H., Ida S., 2010, Exoplanets, ed Seager S.. Tucson,
AZ, University of Arizona Press, 347
Lubow S. H., Martin, R G., 2012, ApJL, 749, 37
Lunine J. I., Stevenson D. J., 1982, Icarus, 52, 14
Martin R. G., Lubow S. H., Pringle J. E., Wyatt M. C., 2007,
MNRAS, 378, 1589
Martin R. G., Lubow S. H., 2011a, MNRAS, 413, 1447
Martin R. G., Lubow S. H., 2011b, ApJ, 740, L6
Martin R. G., Lubow S. H., Livio M., Pringle J. E., 2012a, MN-
RAS, 420, 3139
Martin R. G., Lubow S. H., Livio M., Pringle J. E., 2012b, MN-
RAS, 423, 2718
Matsumura S., Pudritz R. E., 2003, ApJ, 598, 645
Matsumura S., Pudritz R. E., Thommes E. W., 2007, ApJ, 660,
1609
Mosqueira I., Estrada P. R., 2003a, Icarus, 163, 198
Mosqueira I., Estrada P. R., 2003b, Icarus, 163, 232
Ogilvie, G. I., & Lubow, S. H. 2003, ApJ, 587, 398
Paardekooper S. J., Mellema G., 2006, A&A, 459, 17
Papaloizou, J. C. B., Nelson, R. P., & Masset, F. 2001, A&A,
366, 263
Perez-Becker D., Chiang E., 2011, ApJ, 727, 2
Pollack J. B., Lunine J. I., Tittemore W. C., 1991, Origin of the
Uranian satellites. In Uranus, eds J.T. Bergstralh, E.D. Miner,
M.S. Matthews, 469, University of Arizona Press, Tucson.
Rafikov, R. R., 2002, ApJ, 572, 566
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Skilling J., Strong A. W., 1976, A&A, 53, 253
Stevenson D. J., Harris A. W., Lunine J. I., 1986, in Satellites,
eds. J. A. Burns, M. S. Matthews, Tucson, AZ, Univ. Arizona
Press, 39
Toomre A., 1964, ApJ, 139, 1217
Umebayashi T., 1983, Prog. Theor. Phys., 69, 480
Valenti J. A., Basri G., Johns C. M., 1993, AJ, 106, 2024
Ward, W. R. 1991, Lunar and Planetary Institute Science Con-
ference Abstracts, 22, 1463
Ward W. R., 1997, ApJ, 482, L211
Ward W. R., Canup R.M., 2010, AJ, 140, 1168
Yu C., Li H., Li S., Lubow S. H., Lin D. N. C., 2010, ApJ, 712,
198
Zhu Z., Hartmann L., Calve, N., Hernandez J., Muzerolle J.,
Tannirkulam A. K. 2007, ApJ, 669, 483
Zhu Z., Hartmann L., Gammie C., 2009, ApJ, 694, 1045
Zhu Z., Hartmann L., Gammie C. F., Book L. G., Simon J. B.,
Engelhard E., 2010, ApJ, 713, 1142
c(cid:13) 2011 RAS, MNRAS 000, 1 -- 7
|
1205.4164 | 1 | 1205 | 2012-05-18T14:32:07 | On the HU Aquarii planetary system hypothesis | [
"astro-ph.EP"
] | In this work, we investigate the eclipse timing of the polar binary HU Aquarii that has been observed for almost two decades. Recently, Qian et al. attributed large (O-C) deviations between the eclipse ephemeris and observations to a compact system of two massive jovian companions. We improve the Keplerian, kinematic model of the Light Travel Time (LTT) effect and re-analyse the whole currently available data set. We add almost 60 new, yet unpublished, mostly precision light curves obtained using the time high-resolution photo-polarimeter OPTIMA, as well as photometric observations performed at the MONET/N, PIRATE and TCS telescopes. We determine new mid--egress times with a mean uncertainty at the level of 1 second or better. We claim that because the observations that currently exist in the literature are non-homogeneous with respect to spectral windows (ultraviolet, X-ray, visual, polarimetric mode) and the reported mid--egress measurements errors, they may introduce systematics that affect orbital fits. Indeed, we find that the published data, when taken literally, cannot be explained by any unique solution. Many qualitatively different and best-fit 2-planet configurations, including self-consistent, Newtonian N-body solutions may be able to explain the data. However, using high resolution, precision OPTIMA light curves, we find that the (O-C) deviations are best explained by the presence of a single circumbinary companion orbiting at a distance of ~4.5 AU with a small eccentricity and having ~7 Jupiter-masses. This object could be the next circumbinary planet detected from the ground, similar to the announced companions around close binaries HW Vir, NN Ser, UZ For, DP Leo or SZ Her, and planets of this type around Kepler-16, Kepler-34 and Kepler-35. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 000 -- 000 (0000)
Printed 13 November 2018
(MN LATEX style file v2.2)
On the HU Aquarii planetary system hypothesis
Krzysztof Go´zdziewski1, Ilham Nasiroglu2,3, Aga Słowikowska4, Klaus Beuermann5,
Gottfried Kanbach2, Bartosz Gauza6,7, Andrzej J. Maciejewski4, Robert Schwarz8,
Axel D. Schwope8, Tobias C. Hinse9, Nader Haghighipour10, Vadim Burwitz2,11,
Mariusz Słonina1 and Arne Rau2
1Toru´n Centre for Astronomy, Nicolaus Copernicus University, Gagarin Str. 11, 87-100 Toru´n, Poland
2Max Planck Institute for Extraterrestrial Physics, Giessenbachstrasse, 85748 Garching, Germany
3University of Cukurova, Department of Physics, 01330 Adana, Turkey
4Kepler Institute of Astronomy, University of Zielona G´ora, Lubuska 2, 65-265 Zielona G´ora, Poland
5Universitat Goettingen, Institut fur Astrophysik, Friedrich-Hund-Platz 1, DE 37077 Gottingen, Germany
6Instituto de Astrof´isica de Canarias (IAC), E-38200 La Laguna, Tenerife, Spain
7Dept. Astrof´isica, Universidad de La Laguna (ULL), E-38206 La Laguna, Tenerife, Spain
8Leibniz-Institute fur Astrophysik (AIP), An der Sternwarte 16, 14482 Potsdam, Germany
9Korea Astronomy and Space Science Institute (KASI), Optical Astronomy Research Center, 776 Daedukdae-ro, Daejeon, 305-348, Republic of Korea
10Institute for Astronomy and NASA Astrobiology Institute, University of Hawaii-Manoa, 2680 Woodlawn Drive, Honolulu, HI 96822, USA
11Observatorio Astronomico de Mallorca, 07144 Costitx, Mallorca, Spain
13 November 2018
ABSTRACT
In this work, we investigate the eclipse timing of the polar binary HU Aquarii that has been ob-
served for almost two decades. Recently, Qian et al. attributed large (O-C) deviations between
the eclipse ephemeris and observations to a compact system of two massive jovian compan-
ions. We improve the Keplerian, kinematic model of the Light Travel Time (LTT) effect and
re-analyse the whole currently available data set. We add almost 60 new, yet unpublished,
mostly precision light curves obtained using the time high-resolution photo-polarimeter OP-
TIMA, as well as photometric observations performed at the MONET/N, PIRATE and TCS
telescopes. We determine new mid -- egress times with a mean uncertainty at the level of 1 sec-
ond or better. We claim that because the observations that currently exist in the literature are
non-homogeneous with respect to spectral windows (ultraviolet, X-ray, visual, polarimetric
mode) and the reported mid -- egress measurements errors, they may introduce systematics that
affect orbital fits. Indeed, we find that the published data, when taken literally, cannot be
explained by any unique solution. Many qualitatively different and best-fit 2-planet config-
urations, including self-consistent, Newtonian N-body solutions may be able to explain the
data. However, using high resolution, precision OPTIMA light curves, we find that the (O-C)
deviations are best explained by the presence of a single circumbinary companion orbiting at
a distance of ∼ 4.5 AU with a small eccentricity and having ∼ 7 Jupiter-masses. This object
could be the next circumbinary planet detected from the ground, similar to the announced
companions around close binaries HW Vir, NN Ser, UZ For, DP Leo or SZ Her, and planets
of this type around Kepler-16, Kepler-34 and Kepler-35.
Key words: extrasolar planets -- LTT technique -- N-body problem -- polar -- star: HU Aqr
2
1
0
2
y
a
M
8
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
6
1
4
.
5
0
2
1
:
v
i
X
r
a
1
INTRODUCTION
Magnetic cataclysmic variables (CVs, polars, a.k.a. AM Her stars)
are interacting close binary systems. They consist of a main se-
quence red dwarf secondary filling its Roche lobe, and a strongly
magnetized white dwarf (WD) primary, with typical magnetic field
values of 10 -- 80 MG (Schwope et al. 2001). The strong magnetic
field of the primary interacts with the weaker magnetic field of
c(cid:13) 0000 RAS
the secondary and locks the two stars together. Hence, the syn-
chronously rotating WD spins at the same rate as the orbital mean
motion of the binary. Under the gravitational field of the primary,
material flows from the donor star initially along the binary orbital
plane, and finally is accreted quasi-radially onto the magnetic poles
of the WD. The variable HU Aquarii system (hereafter HU Aqr) be-
longs to this class of CV binaries hosting a strongly magnetic WD
accompanied by a red dwarf (spectral type M4V) with an orbital
2
K. Go´zdziewski et al.
period of about 125 minutes. This system is one of the brightest
polars in the optical domain with visual magnitudes ranging from
14.6 to 18 (Warner 1995; Hellier 2001), as well as in the X-ray
energy range. Therefore, it has also been one of the most studied
systems so far.
Accreted matter leaving from the red dwarf is initially not
affected by the magnetic field of the WD. The matter follows a
ballistic trajectory up to the moment when the WD magnetic field
begins to dominate. Because the WD magnetosphere extends be-
yond the L1 radius, the plasma stream cannot orbit freely, and thus
does not form an accretion disk, unlike in other non-magnetic cat-
aclysmic variables. The accreted matter follows the magnetic field
lines and forms an accretion spot at the magnetic poles of the WD.
In many systems, the WD magnetic field is tilted in a such way that
one magnetic pole is oriented toward the direction of flowing mat-
ter. Eclipses observed in highly inclined polars provide information
about the stream geometry.
According to the most recent work of Schwope et al. (2011)
the inclination of the binary is ∼ 87o ± 0.8o. This special geom-
etry is important for the planetary hypothesis investigated in this
work. Assuming that a planetary companion (or companions) have
formed in the circumbinary disc, the inclination constraint removes
the mass indeterminacy inherent to the eclipse timing method.
Recently, the HU Aqr system has received much attention in
the literature. Schwarz et al. (2009) carried out an analysis of the
light curves of the system and derived mid -- egress times of the po-
lar. They proposed a planetary companion as one possible expla-
nation of the detected (O-C) variability. Shortly after this work,
Qian et al. (2011) presented and discussed 10 new light curves in
the optical domain. These authors confirmed the deviations of the
observed mid -- egress times from a linear or quadratic ephemeris,
concluding that the large (O-C) residuals may be explained by the
Light Travel Time [LTT aka Roemer effect, Irwin (1952)] due to
two jovian-mass planetary companions in orbits with semi-major
axes of a few AU and a moderate eccentricity of ∼ 0.5 for the outer
planet. The orbit of the inner planet was fixed to be circular. The ra-
tio of the orbital periods of these massive putative planets would be
presumably in a low -- order 2c:1b mean motion resonance (MMR).
The latter points to significant mutual interactions between these
objects which strongly affects the orbital stability of the system.
Indeed, shortly after that work was published, Horner et al. (2011)
performed dynamical analysis of the putative HU Aqr 2-planet sys-
tem, exploring the parameter space within 3σ uncertainty levels of
the derived Keplerian elements. They found that none of the best-
fit configurations presented by Qian et al. (2011) were dynamically
stable implying that the planetary hypothesis proposed by these au-
thors is hard to maintain. After a few months, in a new paper, Wit-
tenmyer et al. (2012) also re-analysed data in Qian et al. (2011)
confirming that the 2-planet configuration is mathematically con-
sistent with the observations, but inferred orbits are catastrophically
unstable over a ∼ 103 -- 104 year time -- scale. Furthermore, in a very
recent paper, Hinse et al. (2012) improved the Keplerian fit models
of this system by imposing orbital stability constraints on the ob-
jective function (χ2
ν)1/2. Although these authors were able to find
a stable 2-planet configuration consistent with the linear ephemeris
model, orbital parameters of these planets were relatively distant
from the formal best-fit solution by more than 3σ. Because the re-
sults of extensive dynamical analysis contradict the 2-planet hy-
pothesis, an alternative explanation of the (O-C) diagrams needs to
be considered.
Long term monitoring of HU Aqr shows large variations of
the accretion rate that could be correlated with a migration of the
accretion spot. Taking into account the observed changes of the ac-
cretion geometry during different accretion states, high and inter-
mediate ones, Schwope et al. (2001) estimated the possible time --
shift of eclipses to be on the level of 2 seconds, which is still much
smaller than the deviations between the theoretical ephemeris and
observed mid -- egress moments. These results suggest that the mi-
gration of the accretion spot cannot be responsible for the (O-C)
deviations, and we therefore ruled it out.
The (O-C) variability of HU Aqr could be also attributed to
other complex astrophysical phenomena in the binary, like the Ap-
plegate mechanism and/or magnetic braking discussed by Schwarz
et al. (2009) and Wittenmyer et al. (2012). The timing signal might
also be affected by non-Gaussian red-noise, which is a well known
effect present in the precision photometry of transiting planets and
timing of millisecond pulsars (e.g., Pont et al. 2006; Coles et al.
2011). Hence, it should be stressed that we focus here on the plan-
etary hypothesis, as one of the possible, simple and somehow at-
tractive explanations of the (O-C) variability. We try to solve "the
puzzle" of unstable 2-planet models through a new and independent
analysis of available data, conducted along three basic directions.
The first one relies on the re-analysis of published data, be-
cause we found a few inconsistencies in the literature. Surprisingly,
while in the recent paper, Wittenmyer et al. (2012) take into ac-
count 82 mid -- egress points from Schwarz et al. (2009) and Qian
et al. (2011), this is not the full data set available in the literature
at that time. In fact, 72 egress times published by Schwarz et al.
(2009) extend the data set in Schwope et al. (2001) that included
31 measurements. Although the early data of Schwope et al. (2001)
spanning cycles 0 -- 22478 overlap with measurements in Schwarz
et al. (2009) in the time window covering cycles 1319 -- 60097, they
may be helpful to constrain the best-fit models. Up to now, the full
list of published observations consists of 113 points, including data
in Qian et al. (2011). Yet it is not quite obvious whether Qian et al.
(2011) included measurements in Schwope et al. (2001) in their
analysis. Hinse et al. (2012) considered the full data set available
at that time, but in terms of the linear ephemeris LTT model only.
In this context, a direct comparison of the results in the published
papers is difficult.
The second aspect of our study is a new kinematic model of
the ephemeris that properly approximates orbits of putative com-
panions in multi-body systems (to the lowest possible order in the
masses), as compared to the full N-body model. The kinematic
model used in all cited papers refers back to Keplerian parametri-
sation by Irwin (1952) for the "one companion" case. That model,
though commonly used in the literature (e.g., Lee et al. 2011),
seems nowadays redundant, as it was introduced to quantify a simi-
larity between the LTT and the radial velocity curves, in a particular
reference frame with the origin at the center of the two -- body LTT
orbit (instead of the dynamical barycenter). Indeed, the recent, al-
though short history of modeling precision radial velocities teaches
us that multiple planetary systems should be modeled either us-
ing kinematic formulation in a proper coordinate frame (e.g. Lee
& Peale 2003; Go´zdziewski et al. 2003), or using the most general
and accurate full N-body model (Laughlin & Chambers 2001). The
dynamical stability can be further incorporated as an additional,
implicit observable to the objective function (e.g., Go´zdziewski &
Maciejewski 2001; Go´zdziewski et al. 2008). In this work we are
focused on the kinematic modeling though the self-consistent N-
body approach was also used to analyse the HU Aqr mid -- egress
times (see Appendix). Our results indicate that the Newtonian
model may be required for other systems presumably exhibiting
the LTT effect, indeed.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
The third and, actually, critical direction of our work, is a
careful independent analysis of the significantly extended data set
including already published egress times, and new high-precision
timing of the egresses obtained with the ultra-fast photometer OP-
TIMA (Kanbach et al. 2003, 2008), as well as the MONET/N,
PIRATE and TCS telescopes. We collected almost 60 new egress
times with superior accuracy at the sub-second level. Moreover, we
found that the literature data are non-homogeneous, as they come
from different instruments with different time resolutions, as well
as working in different spectral windows (from the visual range,
through the UV, to the X-ray domain) and non/polarimetric modes.
Taking into account the above mentioned inhomogeneities factors
and new data, we present the results from a quasi-global optimiza-
tion of two basic LTT models, leading us to the conclusion that
the measured (O-C) data of HU Aqr may be best explained by
a 1-planet configuration. Simultaneously, it would resolve the 2-
companions instability paradox in the simplest way.
This paper is structured as follows. In Sect. 2 we derive 2-
planet LTT models on the basis of Jacobi coordinates which de-
scribes kinematic orbits in multiple systems properly, as well as a
hybrid optimisation algorithm and numerical setup that makes it
possible to explore the (χ2
ν)1/2 parameter space in a quasi-global
manner. We also briefly describe the N-body formulation of the
LTT effect. In Sect. 3, we re-analyse the data set published in the
literature, following the 2-planet hypothesis by Qian et al. (2011)
and further investigated by Wittenmyer et al. (2012). Two examples
of highly degenerate best-fit solutions are found. In Sect. 4, possible
effects of different spectral windows for the light curves and deter-
mination of egress times are studied. Furthermore, we describe the
new data set derived with the OPTIMA and other instruments. In
Sect. 5, we propose the 1-planet model that best explains the (O-C)
variability. We briefly discuss the effect of red-noise in Sect. 6 and
present a summary of our work in the Conclusions, Sect. 7. The
Appendix contains extensive supplementary material to Sect. 5, in-
cluding the results of kinematic and N-body modeling of 2-planet
systems, accompanied by the long-term stability tests.
2 LTT MODEL FOR A 2-PLANET SYSTEM
We briefly develop the Keplerian model of the LTT signal in the
three-body configuration, assuming that a compact binary (like
HU Aqr) has two planetary companions. More technical details
and a generalization of that model will be published elsewhere
(Gozdziewski et al., in preparation). We consider the compact bi-
nary as a single object having the mass of m∗, which is reasonable
in accordance with the extremely short orbital period (∼ 125 min)
of the polar. A single companion, as well as multiple-planet models
are particular cases of this problem. The key point is that the Ke-
plerian (or kinematic) model requires special coordinates in order
to preserve the sense of Keplerian elements as an approximation
of the exact N-body initial condition. That can be accomplished by
expressing the dynamics through particular canonical coordinates
in which the mutual planetary interactions are possibly small with
respect to the main, "pure" Keplerian part. The barycentric formu-
lation (Irwin 1952) in fact ignores the interactions which could be
adequate for low-mass circumbinary objects, but it might fail when
they have stellar masses as in the SZ Her system (Lee et al. 2011)
where companions are as massive as 20% of M(cid:12), and can shift the
system barycenter significantly. The reason for introducing this im-
proved model is in fact the same as in the precision radial velocities
analyses (e.g., Lee & Peale 2003; Go´zdziewski et al. 2003).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
3
2.1 Kinematic parametrization of the LTT effect
(1)
m1
m2
κ2 =
m1 + m∗ ,
m1 + m2 + m∗ .
One of the well known frames that provides a proper description
of kinematic orbits in multiple systems is Jacobi coordinates. Let
us assume that m∗, m1 and m2 represent the masses of the com-
pact binary m∗ and two planets, respectively. Let us also assume
that the Cartesian coordinates of these objects with respect to the
three-body barycentre are R∗,R1,R2, and their Jacobi coordinates
are denoted by r∗ ≡ R∗,r1,r2 (see Fig. 1). Here R∗ is the posi-
tion of the centre of mass of the binary (CMB) in the barycentric
frame, and r1, r2 are position vectors of the planetary companions
in the Jacobi frame. In this formalism, the barycentric position of
the binary is:
R∗ = −κ1r1 − κ2r2,
where the mass factor coefficients κ1 ≥ 0, κ2 ≥ 0 are given by:
κ1 =
(2)
The coordinate transformation R → r is taken from Malhotra
(1993):
r∗ ≡ R∗,
r1 = R1 − R∗,
r2 = R2 − m∗R∗ + m1R1
and the inverse transformation is derived from the integral of the
barycentre:
R∗ = −κ1r1 − κ2r2,
R1 = (1− κ1)r1 − κ2r2,
R2 = (1− κ2)r2.
To the first order in the mass-ratio (∼ m1,2/m∗), the true N-body
orbit of body i, (i = 1,2) is described through geometric Keplerian
elements as follows:
ri(t) = Pi [cosEi(t)− ei] + Qi
where
Pi = ai (li cosωi + ki sinωi) , Qi = ai (−li sinωi + ki cosωi) ,
and geometric elements are defined through:
i sinEi(t),
m1 + m∗
1− e2
(cid:113)
(3)
(4)
,
+sinΩi
+cosΩi
,
0
li =
+cosii cosΩi
−cosii sinΩi
.
sinii
ki =
i a3
Here, Ei(t) is the eccentric anomaly derived from the Kepler equa-
tion
ni(t − Ti) = Ei(t)− ei sinEi(t),
where ni = 2π/Pi is the mean motion, in accordance with Kepler
3rd law, n2
i = µi, where Pi is the orbital period of a given object.
Two tuples (ai,ei,ii,Ωi,ωi,Ti), i = 1,2, that consist of the semi-
major axis, eccentricity, inclination, nodal angle, argument of peri-
centre, and the time of pericentre passage, respectively, are for the
geometric Keplerian elements. These are related to the Cartesian
coordinates in the Jacobi frame through the usual two-body formu-
lae (see, e.g. Morbidelli 2002), with an appropriate mass parameter
µi (see below).
From Eq. 1, the Z∗ component of the CMB with respect to the
system barycentre is:
Z∗(t) ≡ R∗ · ez = −κ1z1(t)− κ2z2(t),
(5)
4
K. Go´zdziewski et al.
Figure 1. The geometry of the system. The binary has a total mass m∗
and because of its short orbital period it can be considered as a point-like
object accompanied by planets as point-masses. The origin of the coordinate
system is fixed at the barycentre of the three-body system. The line-of-sight
is along the z-axis. See the text for more details.
where ez is the unit vector along the z -- axis of the reference frame,
directed toward the observer. The signal contribution due to a given
companion is:
zi(t) = ai sinii
sinωi (cosEi(t)− ei) + cosωi
1− e2
i sinEi(t)
(6)
(cid:113)
(cid:20)
(cid:21)
(for planets i = 1,2). The zi(t) are then combined to obtain the Z∗(t)
component of the CMB position vector w.r.t. the system barycentre.
The LTT signal is then expressed as:
τ(t) = − 1
c
(cid:18) m1
Z∗ ≡ +
m1 + m2 + m∗
m1 + m∗
(cid:19)
z1 +
m2
1
c
z2
,
where c is the speed of light. Note that we used the planetary ver-
sion of the three-body system, with one dominant mass (m∗), hence
the gravitational Keplerian parameters are:
µ1 = k2(m1 + m∗),
µ2 = k2 m∗(m1 + m2 + m∗)
m1 + m∗
,
consistent with the expansion of the Hamiltonian perturbation for
the planetary version of the problem (see, e.g., Malhotra 1993), and
the quantity k denotes the Gauss constant.
We introduce the signal semi-amplitude factors, K1 and K2 as:
(cid:18) 1
(cid:18) 1
c
(cid:19) m1
(cid:19)
m1 + m∗
m2
a1 sini1,
c
m1 + m2 + m∗
a2 sini2.
K1 =
K2 =
(7)
(8)
Using Eq. 6, the single-planet signal contributions ζi are then given
by:
ζi(t) = Ki
sinωi (cosEi(t)− ei) + cosωi
1− e2
i sinEi(t)
.
(9)
(cid:113)
(cid:20)
(cid:21)
the
set of
equation,
i = 1,2, similar
free orbital parameters
is
In this
(Ki,Pi,ei,ωi,Ti),
to the common kinematic
radial velocity model. The orbital period Pi and the time of
pericenter passage are introduced indirectly through the time
dependence expressed by Ei(t).
We would like to note here that the contribution of the planet
as expressed in Irwin's model has an extra term ei sinωi that ap-
pears due to the particular choice of the coordinate system with the
origin at the center of the binary orbit around the common center
of mass of the system. It should also be stressed, that no simple
superposition of kinematic orbits does account for the mutual grav-
itational interactions directly, but in our formulation, the Keplerian
elements are the closest to the osculating N-body initial condition
within the kinematic model.
2.2 The (O-C) formulation
From Eq. 9, the fit model of the planetary-induced LTT signal is
τ(t,K1,P1,e1,ω1,T1,K2,P2,e2,ω2,T2) = ζ1(t) + ζ2(t).
Now let us assume that the observational data are given through
eclipse cycle number l (l = 0, . . . ,N), the date of the eclipse time-
mark tl, and its uncertainty σl. Then the l-cycle eclipse ephemeris
with respect to the reference epoch t0 (l = 0), at time t ≡ tl may be
written as follows:
Tep(l) = t0 + lPbin + τ(tl,K1,2,P1,2,e1,2,ω1,2,T1,2) + "physics",
where Pbin is the orbital period of the binary. It should not be as-
sumed as known in advance, hence must be fitted, as well as the ini-
tial epoch t0 corresponding to cycle number l = 0, simultaneously
with other parameters of the model. The term coded as "physics"
contains non-Keplerian effects, such as the period damping or other
phenomena that may/should be included in the fit model. Here,
we introduce two instances of such a model. Following Hilditch
(2001), the linear ephemeris model, as above,
(O-C) = Tep(l)−t0 − lPbin = τ(tl,K1,2,P1,2,e1,2,ω1,2,T1,2),
and the quadratic ephemeris model, the simplest, yet non-trivial
generalization of the polynomial ephemeris (Hilditch 2001)
(O-C) = Tep(l)−t0−lPbin−βl2 = τ(tl,K1,2,P1,2,e1,2,ω1,2,T1,2).(11)
The quantity β in Eq. 11 is a factor that describes the binary pe-
riod damping (change) due to the mass-transfer, magnetic braking,
gravitational radiation, and/or influence of a very distant compan-
ion
(10)
β =
1
2
Pbin Pbin.
Let us note that also β should be fitted simultaneously with other
free parameters of the model. In the rest of this paper, we use a
common notation in the extrasolar planets literature, that enumer-
ates the planets by subsequent letters, i.e., "b" ≡ "1", "c" ≡ "2",
etc., to avoid any confusion.
2.3 Newtonian model of the LTT effect
A derivation of the N -- body model of the LTT is basically very
simple. It requires the computation of the planetary contribution τ
to the (O-C) signal through the numerical integration of the equa-
tions of motion, a computation of the star barycentric vector R∗
and its Z∗ -component, in accord with Eq. 5. This formulation ac-
counts for the mutual interactions between all bodies in the system.
A serious computational drawback of this model is a significant
CPU overhead, nevertheless, as we will show in the Appendix, its
application for systems with massive companions presumably in-
volved in low order mean motion resonances can be indispensable,
To solve the equations of motion efficiently, we used the ODEX2
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Rrmm*mR2211r12zyxto observerR*r*integrator (Hairer et al. 2009) designed for conservative, second or-
der ordinary differential equations (ODEs). The imposed variable
time step accuracy preserved the total energy and the angular mo-
mentum better than 10−11. In terms of the Newtonian model, the
planetary bodies are parametrised through the mass mi, semi-major
axis ai, eccentricity ei and three Keplerian angles describing the
orientation of the orbit, for each companion in the system. We also
assume that the binary is a point mass with the prescribed total
mass of the binary. Assuming a coplanar configuration (in the N --
body model the same inclinations are "absorbed" in the planetary
masses), we have 5 free orbital parameters for each planet, similar
to the kinematic model. Here, they are then represented as "usual"
osculating, astrocentric Keplerian elements at a given initial epoch,
but other types of the osculating elements may be used as well.
2.4 The optimization method and numerical setup
ν)1/2(K1,2,P1,2,e1,2,ω1,2,T1,2,t0,Pbin,β),
Having the egress times measured with a great precision (at the
1 second level, or even better), the next step is to determine the
set of primary parameters of the kinematic model, usually with
the least squares approach, by constructing the reduced (χ2
ν)1/2 --
squared function
ν)1/2 ≡ (χ2
(χ2
and searching for its minimum in the space of the model parame-
ters. It is well known, however, that the (χ2
ν)1/2 function may pos-
sess many local minima, particularly if the model is not well con-
strained, as it might be in our case. To seek a global solution, we
apply a hybrid algorithm that consists of two steps: a quasi-global
method, the Genetic Algorithm [GA, Charbonneau (1995)] that is
relatively slow and inaccurate, but makes it possible to find good
approximations to the second step, a fast local method. Here, we
use the Levenberg-Marquardt (L-M) algorithm with analytically
computed derivatives. The idea of the hybrid code comes from
our earlier works on modelling radial velocity observations (e.g.,
Go´zdziewski & Konacki 2004). We used freely available Fortran
codes of the Genetic Algorithm (PIKAIA1, by Phil Charbonneau &
Barry Knapp) and of the L-M method from the well known MIN-
PACK2 package.
Once the primary set of the orbital model parameters are de-
termined in the form of two five-tuples (Ki,Pi,ei,ωi,Ti), i = 1,2,
we may also derive inferred Keplerian elements, such as minimal
planetary masses and semi-major axes, by solving nonlinear equa-
tions expressing Ki (Eqs. 7, 8) and the 3rd Kepler law in terms of
the primary model parameters. The inclination has to be held fixed.
Hence usually one assumes ii = 90◦. Let us underline that while the
LTT-model (Eq. 9) formulated in the barycenter frame has the same
mathematical form as in the Jacobi frame, the orbital, geometrical
(Keplerian) elements in multiple systems should be related to Jaco-
bian, canonical coordinates. If in the N-body numerical integrations
and stability studies, initial conditions have to be in the form of os-
culating elements, one should transform these Jacobian elements
into the Cartesian coordinates w.r.t. the Jacobi frame (e.g., Mor-
bidelli 2002), and then, if necessary, to the astrocentric or barycen-
tric coordinates. In this sense, "barycentric" and "Jacobian" two-
body elements may closely coincide for small, Jovian-mass plan-
ets. But for more massive companions when the LTT signal is easier
to detect, or for very compact (resonant) systems, the semi-major
1 http://www.hao.ucar.edu/modeling/pikaia/pikaia.php
2 http://www.netlib.org/minpack/
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
5
axes, masses, Keplerian angles and inferred N-body initial condi-
tions may be significantly different in both frames. We will discuss
this issue in more detail in a forthcoming article (Gozdziewski et
al., in preparation).
Each run of the hybrid code has been initiated by random se-
lection of the GA population (between 512 and 4096 individuals),
considering possibly wide parameter ranges. For instance, the range
of orbital periods was set blindly to [800, 63600] days, and angles
and eccentricities were set to their whole possible ranges. The orig-
inal "population" was then transformed by GA operators over 512 --
1024 generations. Each member of the final set was then used as
an initial condition for the L-M algorithm, and the resulting solu-
tions were sorted and stored. The hybrid procedure was repeated
hundreds of times for each combination of model -- data set. We ex-
amined whether the obtained solutions converged to the same min-
ima. Due to the semi-deterministic nature of the GAs, one should
interpret the results in a statistical sense.
The same procedure may be applied to the Newtonian model,
as the planetary contribution τ can be computed independently of
the optimisation method. (In this case the derivatives to the LM
algorithm were approximated numerically).
Finally, uncertainties of the best-fit parameters were deter-
mined using the bootstrap algorithm (Press 2002), as the variances
of parameters in a tested solution that has been re-fitted to 4096 syn-
thetic data sets drawn randomly with replacement from the original
sample. We found that due to the particular distribution of OPTIMA
observations that are grouped in small "clumps" of a few data
points, the bootstrap algorithm tends to underestimate the uncer-
tainties when compared to the formal error determination through
the diagonal elements of (χ2
ν)1/2 curvature (covariance) matrix.
3 KINEMATIC MODELING THE LITERATURE DATA
To verify the literature models of the HU Aqr system, we gathered
Barycentric Julian Dated (BJD) egress times published by Schwope
et al. (2001), Schwarz et al. (2009) and Qian et al. (2011). That data
set consists of 113 points, and will be called the SSQ set hereafter.
Our first attempt was to reproduce the results of Qian et al. (2011)
with our formulation of the LTT model. We did not expect this to
be straightforward, since their model assumes the inner planet to be
on a circular orbit. We conducted calculations for two ephemeris
models, linear and quadratic (Eqs. 10 and 11), respectively.
3.1 The linear kinematic ephemeris 2-planet model
In the linear ephemeris case, we found many, almost equally good
ν)1/2 ∼ 1.15 and an rms ∼ 2.3 sec. In
2-planet solutions with (χ2
these best fits the inner planet has a period of ∼ 5500− 6000 days.
However, the period of the outer planet varies between 7000 --
20000 days. The resulting systems imply (O-C) residuals caused in-
dividually by the planets in wide ranges, up to ∼ 6000 seconds, and
companions in basically any mass, eccentricity and period range
while still preserving excellent rms ∼ 2.3 sec and similar "flat" be-
haviour of the residuals. The left panel of Fig. 2 shows the most
exotic and actually the best -- fit solution found in our experiment.
The Keplerian fit parameters of this solution, as well as its inferred
elements are given in Table 1 (Fit A). This fit is very different from
those found by Qian et al. (2011), Wittenmyer et al. (2012), and
even in the last paper by Hinse et al. (2012). This configuration
ν)1/2 ∼ 1.143 and an rms ∼ 2.3 sec, and is characterised by
has (χ2
K. Go´zdziewski et al.
6
almost equal orbital periods of ∼ 5470 days. The pericenter argu-
ments of the planets in this fit differ by nominal value of 180.2◦
and as a result, the Keplerian barycentric orbits are almost exactly
anti-aligned, with planets placed close to their periastrons at the
initial epoch. This configuration could be understood as a pair of
Trojan -- planets in 1c:1b mean motion resonance (MMR). Although
the resulting LTT signal has apparently small amplitude ∼ 60 sec-
onds as shown in the (O-C) diagram (see the left-hand panel in
Fig. 2), the LTT semi-amplitudes Kb,c are excessively large (up to
∼ 6000 seconds), implying just absurdly massive companions of
∼ 10,000 Jupiter mass each (∼10 M(cid:12) !). This solution reveals that
an inherent degeneracy of the LTT signal (and its model) may ap-
pear because the signal is the result of the differential gravitational
tugs of the companions on the binary. Indeed, in this particular Tro-
jan configuration, even small deviations from the anti-alignment of
orbits leads to large changes in the planetary masses (over 3 orders
of magnitude) and semi-major axes (within a range of a few AU),
indicating that as they are not supported by the currently available
observations, these dynamical parameters are poorly constrained.
The mathematical fit permits putative companions as massive as
stars but in reality, such objects should influence dynamical and
spectral properties of the binary system. Such solutions are there-
fore excluded.
The 1c:1b MMR solution is a vivid example demonstrating
that due to the possibility of configurations involved in extremely
strong mutual interactions, modeling the LTT signal globally (with-
out any a priori assumptions on the system configuration) cannot
be studied in terms of the kinematic model. In general, an exact,
self-consistent N-body model should be used to determine the ini-
tial conditions.
3.2 The quadratic kinematic ephemeris 2-planet model
In the case of a quadratic ephemeris, we found a well defined min-
ν)1/2 ∼ 0.972, which is an apparently statistically per-
imum of (χ2
fect solution. Its synthetic curve with measurements over-plotted
is shown in the right -- hand panel of Fig. 2, and orbital parameters
are given in Table 1 as Fit B. That solution has been frequently
found in different runs of the hybrid code, which reinforces its
global character. To show the latter, we computed parameter scans
of (χ2
ν)1/2 in the (Pb,c,eb,c) -- planes (Fig. 3), by fixing points of a
grid in a given plane and minimizing (χ2
ν)1/2 over all remaining
free parameters of the model. This made it possible to obtain stan-
dard confidence levels as marked with coloured curves. The best-fit
solution is again very different from solutions found in the litera-
ture. While the elements of the inner planet are well constrained,
the orbit of the outermost companion reveals extremely large ec-
centricity (∼ 1). That points again to highly degenerate (unrealistic)
best-fit solutions, with near-parabolic or even hyperbolic, open or-
bit of one "planet" -- as the fit implies -- being a low-mass stellar
object of ∼ 160 Jupiter-masses. Other solutions with slightly worse
ν)1/2 ∼ 1.1 and still very similar rms ∼ 2.2 sec may be found too,
(χ2
which means that the quadratic ephemeris model is unconstrained
by the SSQ data.
In the quadratic ephemeris model, the orbital periods are close
to the 4:3 ratio, which is equivalent to the low-order 4c:3b mean
motion resonance. In addition, the eccentricity of the outer planet is
extreme, close to 1. Hence again, the kinematic formulation seems
inadequate to derive the proper initial condition of the multiple-
planet configuration. We conclude here that when we only have the
SSQ data at our disposal, there seems to be no unique and phys-
Table 1. Jacobian geometric parameters with the inferred masses and semi-
major axes of 2-planet LTT fit models in the barycenter frame with the
linear and parabolic ephemeris to the SSQ data set. Synthetic curves with
data sets are shown in two panels of Fig. 2 and (χ2
ν)1/2 scans in Fig. 3.
Numbers in parentheses represent the uncertainties at the last significant
digit. Total mass of the binary is 1.08 M(cid:12) (Schwarz et al. 2009). Indices
"b" and "c" refer here to planets "1" and "2" in the mathematical model
Eqs. 10 and 11. See the text for more details.
Model
parameter
Fit A
Fit B
linear ephemeris
parabolic ephemeris
Kb [seconds]
Pb [days]
eb
ec
ωb [degrees]
Tb [BJD 2,440,000+]
Kc [seconds]
Pc [days]
ωc [degrees]
Tc [BJD 2,440,000+]
Pbin [days]
T0 [BJD 2,440,000+]
β [×10−13 day cycle−2]
mb sini [MJup]
ab [au]
ac [au]
mc sini [MJup]
N
(χ2
ν)1/2
rms [seconds]
5928 ± 15
5467 ± 418
0.138 ± 0.034
207 ± 21
11694 ± 175
5942 ± 17
5476 ± 424
0.141 ± 0.035
27 ± 20
6214 ± 451
0.086820400(4)
9102.92004(2)
--
1.375
9780
1.374
9811
113
1.143
2.31
10.2 ± 0.5
2910 ± 28
0.34 ± 0.07
22 ± 8
5652 ± 97
322 ± 30
3931 ± 50
0.99 ± 0.05
358.3 ± 0.2
9207 ± 42
0.0868204250(8)
9102.91988(2)
-3.06(6)
4.08
5.69
4.58
159
113
0.972
2.13
Figure 4. OPTIMA hexagonal fibre bundle centred on HU Aqr. The ring
fibres (1 -- 6) are used to monitor the background sky simultaneously.
ically meaningful solution explaining the LTT variability. Or, the
planetary fit model and its assumptions are incorrect.
4 NEW OBSERVATIONS AND DATA REDUCTION
4.1 Observations with OPTIMA and other instruments
To resolve the model degeneracies as described above, we gathered
new, yet unpublished observations of the HU Aqr binary. The new
collected mid -- egress BJD times are given in Table 2. These data ex-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
7
Figure 2. Synthetic curves of the best-fit, 2-planet solutions to the mid -- egress BJD times of the SSQ data set. The left panel corresponds to a linear ephemeris
model (Eq. 10) and the right panel corresponds to a quadratic (parabolic) ephemeris model (Eq. 11, the parabolic trend has been removed). Panels are labeled
with the orbital periods and eccentricities of the putative companions. Bottom plots with shaded background show the residuals after subtracting planetary and
astrophysical contributions from the LTT signal. Discontinuous-like features of the parabolic ephemeris model around cycles 0 and 46,000 appear due to the
extreme eccentricity of the outer body. See Table 1 for the orbital and inferred elements (Fit A and B, respectively).
Figure 3. Parameter scans of the (χ2
panel in Fig. 2 and Fit B in Table 1). Colour curves are for the formal 1,2,3σ -- levels of the best-fit solution whose parameters are marked with an asterisk.
ν)1/2 function computed around the best fit solution to the SSQ data for the quadratic ephemeris model (see the right-hand
tend the work of Schwope et al. (2001), Schwarz et al. (2009) and
Qian et al. (2011). The currently available data set of HU Aqr egress
times consists of 171 measurements in total, including 10 points
presented in Qian et al. (2011). Among these measurements, 68
were obtained with the Optical Timing Analyzer (OPTIMA) in-
strument that operates mostly at the 1.3 m telescope at Skinakas
Observatory, Crete, Greece.
The high -- speed photometer OPTIMA is a sensitive, portable
detector to observe extremely faint optical pulsars and other highly
variable astrophysical sources. The detector contains eight fibre --
fed single photon counters -- avalanche photodiodes (APDs), and
a GPS for the time control. There are seven fibres in the bundle
(Fig. 4) and one separate fibre located at a distance of ∼ 1(cid:48). Sin-
gle photons are recorded simultaneously and separately in all chan-
nels with absolute UTC time -- scale tagging accuracy of ∼ 4 µs.
The quantum efficiency of the APDs reaches a maximum of 60%
at 750 nm and lies above 20% in the range 450 -- 950 nm (Kan-
bach et al. 2003, 2008). During the HU Aqr observations, OP-
TIMA was pointed at RA(J2000) = 21h07m58.s19, Dec(J2000) =
−05◦17(cid:48)40.(cid:48)(cid:48)5, corresponding to the central aperture of the fibres
bundle (Fig. 4). For sky background monitoring, we usually choose
one out of the six hexagonally located fibres. We look for the fibre
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
that is not by chance pointed to any source, therefore records sky
background, and its response is the most similar to the central fibre
response when the instrument is targeted at the dark sky. An exam-
ple of a sky background subtracted light curve is shown at the top
left-hand panel in Fig. 5.
We derived new fits to the HU Aqr eclipse egress times, as
well as reanalysed many of the already published OPTIMA data.
There are 26 eclipses obtained by OPTIMA published by Schwarz
et al. (2009). We were able to reanalyse only 21 out of the 26
light curves, because only those were available in the OPTIMA
archive. Our completely new data set includes 42 precision photo-
metric observations, starting from cycle l ∼ 29,900, overlapping in
time window with the literature data. We derived 23 new eclipse
profiles from the OPTIMA data archive spanning 1999 -- 2007 and
obtained 19 new OPTIMA optical HU Aqr light curves in 2008 --
2010. Note that some of the OPTIMA observations have been al-
ready published in the very recent literature (Nasiroglu et al. 2010).
We also gathered and reduced 11 observations performed at
the MONET (MOnitoring NEtwork of Telescopes) project which is
a network of two 1.2 m telescopes operated by the Georg-August-
Universitat Gottingen, the McDonald Observatory, and the South
African Astronomical Observatory. These precision data in white
-50-40-30-20-10 0 10 20O-C [seconds]HU Aqr 2-planet LTT-fit (linear ephemeris, SSQ data), √χ2r=1.143, rms=2.3 sPb = 5467 ± 418 days eb = 0.14 ±0.04Pc = 5476 ± 424 days ec = 0.14 ±0.04 -5 0 5 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-10 0 10 20 30O-C [seconds]HU Aqr 2-planet LTT-fit (quadratic ephemeris, SSQ data), √χ2r=0.972, rms=2.13 sPc = 3931 ± 25 daysec = 0.999 ± 0.05Pb = 2910 ± 28 dayseb = 0.34 ± 0.05-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]28002850290029503000orbital period Pb [days]0.10.20.30.40.50.6eccentricity eb0.960.9811.02√χ2r1.06❂38003850390039504000orbital period Pc [days]0.60.70.80.91.0eccentricity ec0.960.9811.02√χ2r1.06❂K. Go´zdziewski et al.
8
light (500 -- 800 nm) were binned in 5 s intervals, with 10−6 days
(0.1 s) accuracy, separated by 3 s readout. The most recent obser-
vations were performed at November 18th, 2011.
An additional three egress times were obtained from the
eclipse observations carried out with the PIRATE telescope
equipped with the SBIG STL1001E CCD camera (Holmes et al.
2011). PIRATE, funded by the Open University Department of
Physics and Astronomy, is a remote-controlled telescope located
at the Astronomical Observatory of Mallorca (OAM), Spain.
We also performed optical observations of HU Aqr in white
light with the 1.5-m Carlos S´anchez Telescope (TCS) equipped
with Wide FastCam (WFC, Fig. 6). The WFC is a 1k × 1k -- pixel
camera with optics offering a FOV of 12 arcmin with a scale of 0.6
arcsec/pix. HU Aqr eclipses were observed on September 30, and
October 01, 2011 with integration times of 3 and 5 seconds, respec-
tively. WFC works in frame transfer mode, therefore the readout
time is effectively null or in other words is equivalent to the expo-
sure time. UTC mid-exposure times of the photometric measure-
ments were converted to the Barycentric Julian Dates in Barycen-
tric Dynamical Time using the procedure developed by Eastman
et al. (2010).
Some technical details of the observations performed with the
MONET/N, TCS, and PIRATE telescopes are given in Table 3.
4.2 Determining time markers of the eclipses
In Fig. 5 we show an example of HU Aqr high time resolution
OPTIMA photometric (see the right top panel, and blue curve in
the left-hand top panel) and polarimetric (Stokes I, red curve in the
right-hand top panel) light curves. These graphs are to be compared
with the light curves from TCS, obtained with 3 and 5 seconds
exposures illustrated in Fig. 6. Obviously, the OPTIMA resolution
makes it possible to track the egress phase closely, which enabled
us to determine the mid -- egress moments very precisely.
Measuring the time of mid -- egress properly is critical to ob-
tain the (O-C) diagrams, since it is the time marker of the eclipse
(Schwope et al. 2001; Schwarz et al. 2009). To derive the mid --
egress moment t0, the sigmoid function
(a2 − a1)
,
(12)
I(t) = a1 +
(1.0 + exp([t0 −t]/∆t)
parametrised by a1,a2 and t0 was fitted to the light curve points
in the egress phase of the eclipse, spanning preselected exponential
scaling parameters ∆t . We found that there is no strong dependence
of the derived t0 on the adopted ∆t. This can be seen in the bot-
tom left-hand panel of Fig. 5 where three mid -- egress times t0 are
marked with black open squares. These moments are derived for
three different choices of ∆t: 0.1, 0.5 and 1.0, respectively. While
these times depend on ∆t, they fall within a 2 second range, as
marked by a shaded strip at the bottom left-hand panel of Fig. 5. A
half of that range (∼ 1 second) may be typically estimated as the
maximum possible error of t0 in the OPTIMA data set. The formal
1σ uncertainty of the sigmoid fit in this case is still smaller and at
the level of ∼ 0.1 second. Moreover, the shape of the eclipse may
significantly depend on the spectral window. Panels in the right col-
umn of Fig. 5 illustrate the light curves of HU Aqr derived in the
optical, white band domain (the blue curve) and in the polarimet-
ric domain (Stokes I, the red curve). In the latter case, the egress
looks quite different and spans over a longer time. Given that these
two data sets were taken four years apart, the observed difference
might have been caused by different emission states of the source.
Table 2. New HU Aqr BJD mid -- egress times on the basis of light curves
collected with: OPT-ESO22 -- OPTIMA photometer installed at ESO
(Chile), OPT-SKO -- OPTIMA operated at the Skinakas Observatory
(Crete), PIRATE -- a telescope at the Astronomical Obs. of Mallorca,
MONET/N -- the network of telescopes at the McDonald Obs. and the
SAO (South Africa), and WFC -- the 1.5-m TCS (Canary Islands).
Cycle
29946
29957
29958
30265
30287
30299
30300
30310
30311
35469
38098
42486
42487
44534
44557
51020
51066
51067
55535
55627
55661
55719
60085
64657
64885
64886
65265
67791
67917
67918
68009
72099
72110
72121
72133
72225
72237
72248
72305
72351
72352
72421
73409
73559
73560
75467
75812
76721
77031
77066
77067
77078
77546
77557
77789
77802
77823
78100
BJD
Error [days]
Instrument
2451702.8443352
2451703.7993545
2451703.8861705
2451730.5400324
2451732.4500902
2451733.4919357
2451733.5787554
2451734.4469740
2451734.5337856
2452182.3533852
2452410.6041626
2452791.5719484
2452791.6587715
2452969.3800760
2452971.3769377
2453532.4971595
2453536.4909030
2453536.5777278
2453924.4913426
2453932.4788164
2453935.4307071
2453940.4662754
2454319.5242409
2454716.4671496
2454736.2622085
2454736.3490181
2454769.2539926
2454988.5622710
2454999.5016391
2454999.5884526
2455007.4891162
2455362.5844371
2455363.5394546
2455364.4944885
2455365.5363444
2455373.5238044
2455374.5656456
2455375.5206715
2455380.4694292
2455384.4631748
2455384.5499944
2455390.5406108
2455476.3190971
2455489.3421698
2455489.4290151
2455654.9954277
2455684.9484608
2455763.8681410
2455790.7824571
2455793.8211556
2455793.9079841
2455794.8630179
2455835.4949490
2455836.4499905
2455856.5922852
2455857.7209399
2455859.5441786
2455883.5934038
0.0000037
0.0000038
0.0000034
0.0000041
0.0000023
0.0000033
0.0000054
0.0000031
0.0000018
0.0000030
0.0000084
0.0000015
0.0000024
0.0000033
0.0000085
0.0000100
0.0000064
0.0000033
0.0000102
0.0000061
0.0000064
0.0000162
0.0000074
0.0000053
0.0000038
0.0000016
0.0000023
0.0000029
0.0000017
0.0000054
0.0000018
0.0000032
0.0000019
0.0000029
0.0000015
0.0000048
0.0000040
0.0000040
0.0000030
0.0000024
0.0000022
0.0000013
0.0000578
0.0000578
0.0001156
0.0000040
0.0000023
0.0000035
0.0000039
0.0000077
0.0000055
0.0000065
0.0000179
0.0000295
0.0000038
0.0000090
0.0000066
0.0000022
OPT-ESO22
OPT-ESO22
OPT-ESO22
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SAO
OPT-SAO
OPT-NOT
OPT-NOT
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
OPT-SKO
PIRATE
PIRATE
PIRATE
MONET/N
MONET/N
MONET/N
MONET/N
MONET/N
MONET/N
MONET/N
WFC
WFC
MONET/N
MONET/N
MONET/N
MONET/N
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
9
Figure 5. High time resolved OPTIMA light curves of HU Aqr. The upper left panel: photometric eclipse in the white light. The bottom left panel: a close-up
around the mid -- egress overplotted with three fitted sigmoid functions with different values set for the ∆t parameter (0.1, 0.5, 1.0, respectively). The fitted mid --
egress times are denoted by open squares. The shaded region corresponds to a time span of 2 seconds. The upper right panel: a comparison of photometric and
polarimetric OPTIMA light curves of HU Aqr obtained in 2004 and 2008, respectively. The bottom right panel: a close-up of photometric and polarimetric
HU Aqr light curves around the mid -- egress time showing the difference between the egress shapes. The shaded region covers 6 seconds.
Table 3. Technical data of the MONET/N, PIRATE and TCS observations
of HU Aqr. Tobs represents the observation time span, and ∆T is the mean
exposure time of a single frame.
Date
Instrument
Filter
Tobs [hours]
PIRATE
PIRATE
TCS/WFC
TCS/WFC
WL
2010 Oct 06
WL
2010 Oct 19
WL
2011 Sep 30
2011 Oct 01
WL
2011 Apr 03 MONET/N WL
2011 May 03 MONET/N WL
2011 Jul 21
MONET/N WL
2011 Aug 17 MONET/N WL
2011 Aug 20 MONET/N WL
2011 Aug 20 MONET/N WL
2011 Aug 21 MONET/N WL
MONET/N WL
2011 Oct 22
MONET/N WL
2011 Oct 23
2011 Oct 25
MONET/N WL
2011 Nov 18 MONET/N WL
1
3.5
1
1
0.60
0.50
0.32
0.60
0.46
0.10
0.58
0.25
0.50
0.16
0.45
∆T [s]
10
10, 20
3
5
8
8
8
8
8
8
8
8
8
8
8
To derive the mid-egress moments gathered in Table 2, the sigmoid
function was fitted to all single light curves.
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
4.3 On the light curves in different spectral windows
Vogel et al. (2008) and Vogel (2008) obtained high time resolved
and accurate light curves of HU Aqr during its low state using the
ULTRACAM (Beard et al. 2002; Dhillon et al. 2007) at the Very
Large Telescope (VLT, May 13, 2005). These authors decomposed
the light curve into three emission components emerging from the
accretion spot, the photosphere surrounding it, and from the white
dwarf itself. As a result, they were able to derive the temperature of
the WD ∼ 13500(200) K and the temperature of the accretion spot
∼ 25500(1500) K. They also estimated the ratio of the spot area to
the WD surface to be on the level of 5%. The black body spectra of
the WD and of the spot have their maxima at 215 nm and 113 nm,
respectively. The accretion spot and the accretion stream are time
variable in brightness, as well as in the geometric position in the
system. Therefore, the orbital phase at which they occur is not con-
stant. The ULTRACAM delivers simultaneous light curves in three
colours: u, g and r. An example of such a three colour HU Aqr light
curve can be found in Fig. 6 of Schwarz et al. (2009), where the
shape -- energy dependence can be easily seen. Thus, a comparison
of egress times in different wavelength domains was possible. In
the two cases with filters u and r, the WD constitutes the main con-
tribution to the egress intensity, because it can be seen unperturbed
when it comes out of eclipse. However, during high and intermedi-
ate accretion states, the WD might be out-shined by the accretion
stream. In the u -- band, the spot contributes 25% of the emission,
300 400 500 600 700 800 900 1000 1100 1200 1300 2800 3000 3200 3400 3600 3800 4000 4200counts rate [cts/s]time [s] 500 600 700 800 900 1000 1100 3710 3720 3730 3740 3750 3760counts rate [cts/s]time [seconds]Sigmoidal fun. with ∆t=1.0Sigmoidal fun. with ∆t=0.5Sigmoidal fun. with ∆t=0.1Fitted mid egress timesOPTIMA data 0 2000 4000 6000 8000 10000-1200-1000-800-600-400-200 0 200 400counts rate [cts/s]time [seconds]OPTIMA white lightOPTIMA Stokes IOPTIMA white light data Jul 22th, 2004OPTIMA Stokes I data Oct 29th, 2008 0 1000 2000 3000 4000 5000 6000-50-40-30-20-10 0 10 20 30 40 50counts rate [cts/s]time [seconds]WL, Jul 22th, 2004Stokes I, Oct 29th, 200810
K. Go´zdziewski et al.
Figure 6. HU Aqr eclipse light curves obtained with the WFC mounted on the TCS. Observations were performed on 30/09/2011 and 01/10/2011 with
integration times of 3 and 5 seconds, respectively. Bottom panels are for the close-up of eclipse egress, with fitted sigmoid function (solid line, see Eq. 12).
Blue vertical lines mark the determined mid -- egress times.
while in the r -- band its contribution is only 12% (Vogel et al. 2008;
Vogel 2008). This suggests that as the time marker of the eclipse,
it is better to use more "reddish" than "blueish" data, particularly
in our case as we have broadband observations gathered in X-rays,
UV and optical domains at our disposal.
There exists evidence that the EUVE light curves differ from
quasi-simultaneous ROSAT/HRI light curves, as it can be seen, for
example, for eclipses recorded on October 1996 and May 1997, as
shown in Fig. 2 of Schwope et al. (2001). The eclipse ingress is
often not measured because of strong suppression of soft X-rays
by absorbing matter along the accretion stream. When the eclipse
duration can be determined, the eclipse duration seems shorter in
the case of EUVE data. Thus the derived mid -- egress moments can
be shifted by a few seconds. According to Schwope et al. (2001),
an expected variation of the eclipse span should be not more than
0.001 of the orbital phase which corresponds to not more than ∼ 8
seconds. Also Schwope et al. (2004) show in their Fig. 3 evidence
of a different eclipse length as well as phase folded egress shapes
at soft X-rays, HST UV, and high-speed optical photometry with
a multichannel multicolor photometer (the MCCP 2.2 m telescope
at Calar Alto) during the 1993 high state and the 1996 low accre-
tion state. The scatter of the egress times resulting from changes
of the accretion geometry during high and intermediate accretion
states is estimated on the level of 2 seconds (Schwarz et al. 2009).
Large differences between light-curves due to the eclipse of the
accretion stream are also visible in the optical photometric mea-
surements performed in parallel with the ROSAT observations (see
Fig. 3 in Schwope et al. 2001). Some of those light curves were
obtained with rather poor time resolution, e.g. 53 and 12 seconds.
It is worth mentioning that time stamps calculated by Schwope
et al. (2001) and Schwarz et al. (2009) for the photon counting UV
and X -- ray detectors were computed from the mean of the arrival
times of the first three photons after the eclipse, while for the opti-
cal observations, they used the moments of the egress half intensity,
which is common in the literature. Examples of the HU Aqr light
curves obtained by the XMM-Newton EPIC -- PN and Optical Mon-
itor detectors are presented in Figures 2, 3 and 4 of Schwarz et al.
(2009). XMM observations, contrary to the bright state ROSAT ob-
servations, were not resolved at time scales shorter than 2 seconds
due to the low count rates.
We first used all available egress times, archival as well as
new ones, to model the (O-C) diagram. However, given the above-
mentioned arguments, we decided to select only those measure-
ments that were obtained in the white light or photometric V band,
in order to keep the data more uniform and homogeneous. This ap-
proach renders the measurements independent of possible varying
emission regions in different bands. We also decided to skip the
most "suspicious" egress-times at some stage of fitting the orbital
model, which is described below.
We note that the HST observations (three points around l ∼
14,000) were performed with the FOS instrument in the 120 --
250 nm range. These points were also excluded in our further anal-
ysis, falling out of the white light and the V band range.
5 MODELING ALL RECENT DATA
Thanks to the new set of precision OPTIMA mid -- egress measure-
ments, as well as observations performed at PIRATE, TCS and
MONET/N telescopes, we can re-fit planetary models to the whole
set of data up to November 18th, 2011. We fitted the data with the
linear and quadratic ephemeris models (Eqs. 10, 11).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
11
Table 4. Keplerian parameters for the 1-planet LTT fit model with quadratic
ephemeris to all data gathered in this work (Fit I) and to measurements se-
lected in the optical and V-band domain (Fit II). Synthetic curves with mid --
egress times overplotted are shown in Figs. 8 and 9. Numbers in parentheses
represent the uncertainty at the last significant digit. Total mass of the binary
is 0.98 M(cid:12) (Schwope et al. 2011). See the text for more details.
Model
parameter
Fit I
Fit II
all measurements
optical measurements
Kb [seconds]
Pb [days]
eb
ωb [degrees]
Tb [BJD 2,440,000+]
13.9 ± 0.3
3278 ± 28
0.03 ± 0.04
211 ± 40
6233 ± 360
14.7 ± 0.2
3287 ± 19
0.13 ± 0.04
226 ± 10
6361 ± 102
Figure 7. Synthetic curve of the 1-planet LTT model with linear ephemeris
to all available data, including the very recent egress times collected by the
OPTIMA photometer, as well as PIRATE, TCS and MONET/N telescopes.
Open circles are for measurements in Qian et al. (2011).
5.1 Single-planet models to all recent data
At the first attempt, we tested the 1-planet hypothesis. For the linear
ephemeris model, the 1-planet solution is characterized by extreme
eccentricity and displays large residuals and a strong trend present
in the (O-C) diagram (see Fig. 7). This suggests a more general
quadratic model, on which we focus now.
The results derived for the whole set of 171 measurements are
shown in the top panels of Fig. 8. Interestingly, the 1-planet model
fits the data very well in a large part of the time-window between
l = 25,000 and l = 80,000 (see the left-hand panel of Fig. 8). How-
ever, over approximately one fourth of the time-window (l = 0 to
l = 25,000), the data fit the synthetic curve poorly. That can be bet-
ter seen in the close-up of the residuals shown in the top right panel
of Fig. 8. It appears that the residuals follow a regular and char-
acteristic "damping" trend, that could be associated with a mass-
transfer process ongoing in the binary or solar-like magnetic cy-
cles. Results of our experiments show that the recent observations
by Qian et al. (2011) appear to be outliers to our 1-planet solution,
as the mid -- egress times are shifted by about of 3 -- 10 seconds w.r.t.
the synthetic curve. Because these observations overlap in the time
window with much more precise OPTIMA data, that discrepancy
between these two data sets cannot be avoided. Actually, obser-
vations by Qian et al. (2011) do not fit any model that has been
tested with the OPTIMA observations, including 2-planet models
and both types of the ephemeris (see the Appendix).
In an effort to explain the strange behaviour of the residuals,
we realized, as it was discussed already, that the available observa-
tions come from different telescopes/instrumentation, and to make
the matter worse, the egress times are measured on the basis of
light curves in different spectral windows. In particular, the first
part of the data set contains the egress times derived from X-rays
(ROSAT and XMM) and ultraviolet (EUVE, XMM OM-UVM2
and HST/FOS) light curves, and some eclipses were observed with
OPTIMA in polarimetric mode. To remove the possible inconsis-
tency due to the different spectral windows and filters, we consid-
ered data sets consisting of the egress times measured only in the
optical range (white light and the V band). The results are shown
in the bottom panels of Fig. 8 for the optical data without X-ray
and UV, but including polarimetric measurements (note, that the
polarimetry was done in the white-light band; compare with the
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
Pbin [days]
T0 [BJD 2,440,000+]
β [×10−13 day cycle−2]
0.0868204226(5)
9102.92004(2)
-2.61 (5)
0.0868204259(4)
9102.91994(1)
-2.95(4)
ab [au]
mb sini [MJup]
N data
(χ2
ν)1/2
rms [seconds]
4.29
6.71
171
5.23
4.8
4.30
7.10
115
2.48
3.7
top panels of Fig. 8 for all data gathered). As can be seen from the
bottom panels in Fig. 8, the "damping" effect has almost vanished,
suggesting that it could have appeared due to the presence of X-ray
and UV-derived eclipses. Still, there is a group of data points with
large errors, around l ∼ 14,000, which do not fit well to the clear
quasi-sinusoidal variation of the (O-C). The deviations of these
points may be explained by poor time-resolution (∼ 12 seconds of
the AIP07 CCD camera), that has been used to observe the HU Aqr
eclipses (Schwope et al. 2001). Let us also note that the Qian et al.
(2011) data points are again systematically outliers with respect to
the synthetic signal. After removing these data and all points (seven
measurements) in the polarimetric mode, we obtained a homoge-
neous optical data set to which we fitted the quadratic ephemeris 1-
planet model again. The synthetic curve of this fit with data points
over-plotted is shown in Fig. 9. Parameters of this fit are presented
in Table 4 as the final solution Fit II and are well constrained by
the observations. To demonstrate the latter, we show projections of
(χ2
ν)1/2 in selected two-dimensional parameter -- planes (see Fig. 10)
close to the best -- fit model. As can be seen, there is a strong corre-
lation between the time and argument of pericentre which can be
understood noting that the orbital phase (λb = ϖb + Mb ) must be
preserved.
The best-fit model seem to constrain the damping factor β ∼
−3 × 10−13 day cycle−2 very well. Such a value is close to es-
timates in the literature, e.g., ∼ −5 × 10−13 day cycle−2 by
Schwarz et al. (2009) and ∼ −2.5 × 10−13 day cycle−2 by Qian
et al. (2011). It is still larger by more than one order of mag-
nitude to be explained by gravitational radiation, but remains in
the range of magnetic braking (Schwarz et al. 2009). A similar
large-magnitude period decrease has been found in other CVs, like
NN Ser (∼ −6× 10−13 day cycle−2, Brinkworth et al. 2006). Be-
sides the angular momentum loss, the large magnitude of the period
change is commonly explained as due to the Applegate mechanism
(basically excluded in the HU Aqr) and/or a the presence of a very
distant, long-period companion body. Likely, a few astrophysical
and/or dynamical effects may be involved that could determine ap-
-25-15 -5 5 15 25 35O-C [seconds]HU Aqr 1-planet LTT-fit (linear ephemeris, all data), √χ2r=18.4, rms=12.6 sPb = 2275 ± 20 dayseb = 0.99 ± 0.01 -40 0 40 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]12
K. Go´zdziewski et al.
Figure 8. The top row: synthetic curve of the 1-planet LTT model with quadratic ephemeris to all available data, gathered in this work, including the very
recent mid -- egress times collected by the OPTIMA photometer, as well as PIRATE, TCS and MONET/N telescopes (the top left) with orbital parameters given
in Table 4 (Fit I), and close-up of residuals to that model (the top right panel). The bottom row: the same for the white light and visual band (V) data, including
polarimetric observations by OPTIMA (i.e., the UV- and X-band observations are excluded) shown at the bottom left panel, and its residuals (the bottom right
panel). The white filled circles mark the Qian et al. (2011) measurements.
Figure 9. Synthetic curve of the 1-planet LTT model with quadratic ephemeris to observations in white light + V band (see the text for more details) without
measurements in Qian et. al (2011) and polarimetric data. Orbital parameters of this solution are given in the Table 4 (Fit II).
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
-15-10 -5 0 5 10 15 20O-C [seconds]HU Aqr 1-planet LTT-fit (quadratic ephemeris, all data), √χ2r=5.23, rms=4.8 sPb = 3278 ± 27 dayseb = 0.03 ± 0.04 -16 0 16 0 10 20 30 40 50 60 70 80resudials [s]cycle number [•103]-15-10-5 0 5 10 15 20 0 10 20 30 40 50 60 70 80residuals [seconds]cycle number [•103]Residuals to 1-planet LTT fit of HU Aqr (quadratic ephemeris, all data)-15-10 -5 0 5 10 15 20O-C [seconds]HU Aqr 1-planet LTT-fit (quadratic ephemeris, visual+polarim.), √χ2r=2.85, rms=4.1 sPb = 3263 ± 19 dayseb = 0.10 ± 0.02 -16 0 16 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-15-10-5 0 5 10 15 20 0 10 20 30 40 50 60 70 80residuals [seconds]cycle number [•103]Residuals to 1-planet LTT fit of HU Aqr (quadratic ephemeris, visual+polarim.)-15-10 -5 0 5 10 15 20 25O-C [seconds]HU Aqr 1-planet LTT-fit (quadratic eph., optical w/o polarim.), √χ2r=2.59, rms=4.1 sPb = 3291 ± 25 dayseb = 0.14 ± 0.04 -20 0 20 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]On the HU Aquarii planetary system hypothesis
13
Figure 10. Colour-coded parameter scans of (χ2
ν)1/2 around the best-fit 1-planet model, to quadratic ephemeris and the optical measurements (Fit II in Table 4).
Its synthetic curve with data points overplotted is shown in Fig. 9. The large symbol marks nominal elements of the solution. Closed curves are for formal
1,2,3σ levels of (χ2
ν)1/2 with scale displayed at the colour legend.
parently secular period decrease. Its definite explanation is com-
plex, and we consider this as a subject of a new, forthcoming work.
We also fitted the quadratic ephemeris model only to the high-
est precision OPTIMA data. The results for measurements that in-
clude polarimetric observations are shown in Fig. 11. For that case,
we found a period similar to the quadratic ephemeris model for the
entire data set. The fit has very small rms ∼ 0.8 sec. The relatively
ν)1/2 ∼ 3.4 of the OPTIMA solution in this case may sug-
large (χ2
gest that the adopted uncertainties, at the ∼ 0.1 -- 0.2 second level
(in a large sub-set of the measurements) are in fact underestimated.
We also identified the most deviating points as coming from the
polarimetric measurements (see e.g., a point marked in the resid-
uals plot around l = 65,000, and the residuals of both solutions).
To examine, whether these data may change the solution, we fitted
the quadratic ephemeris model to the white light OPTIMA obser-
vations only, skipping all polarimetric data. The best-fit orbital pe-
riod of ∼ 3400 days remains close to the full-coverage window fit.
A slightly smaller rms of ∼ 0.7 seconds suggest a better fit with-
out the polarimetric data, indeed. The orbital periods coincidence
cannot be fully proved due to the relatively narrow observational
window of the OPTIMA white light measurements. Actually, the
parameter scan (not shown here) reveals that the 1σ contour around
the (χ2
ν)1/2 minimum in the ( Pb,eb) -- plane is "opened" on the right
side of the orbital period -- axis, hence it can not be constrained yet
by the OPTIMA data alone.
5.2 Alternate models to all recent data
Finally, using the hybrid optimiser, we performed additional ex-
periments by fitting three models to all available data: the 1-planet
model with a heuristic sine damping term, and 2-planet models,
both in terms of the linear and quadratic ephemeris. We also per-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
formed N -body modeling of the 2-planet configurations. The re-
sults, which are described in the Appendix, imply that all these
models lead to non-unique solutions or configurations with simi-
lar orbital periods, for which the kinematic model is inadequate, as
we discussed above. Some of these kinematic best -- fit 2-planet so-
lutions are qualitatively similar to configurations found for the SSQ
data set, with orbital periods ratios close to 1c:1b and 4c:3b, respec-
tively. The extended data set still does not constrain the Keplerian
2-planet models.
The same can be concluded for the N-body models (see Ap-
pendix, Sect. A3). Although we found stable configurations in
terms of the quadratic ephemeris, the semi-major axis of the outer
companion is unconstrained (between 4 AU and at least 20 AU).
These stable fits exhibit varying sign of β (it means that the binary
period might decrease or increase). Moreover, stable solutions with
relativelty small β may be found in very narrow stability zones in
the (ac,ec) -- plane, see the Appendix and Fig. A6. These areas are
associated with low order MMRs, like 3c:2b MMR. It is very uncer-
tain though, how massive companions of HU Aqr could be locked
in such tiny stability areas. Hence, some larger values of ac provid-
ing extended zones of stable motions seems more likely (see panels
of stable fits labeled by IV, V, and VI of Fig. A6). However, there is
also a correlation between the magnitude of β and the semi-major
axis of the outer planet. For relatively distant planet c, β may be
∼ 2× 10−12 day cycle−2, which is difficult to explain by magnetic
braking or mass -- loss. However, this may indicate a presence of a
third companion in an unconstrained orbit.
We conclude that these results seem to favour the 1-planet
hypothesis as the simplest model explaining the (O-C) variability,
particularly in the light of very small rms of the homogeneous OP-
TIMA set and apparently perfect quasi-sinusoidal fit illustrated in
Fig. 11.
14.414.514.614.714.814.915.015.1semi-amplitude Kb [seconds]-3.10-3.05-3.00-2.95-2.90-2.85-2.80damping factor β [10-13 days cycle-2]2.452.502.552.602.652.702.75❂50100150200250300350pericenter passage Tb [BJD +2,456,000 days]205210215220225230235240245argument of pericenter ωb [deg]2.452.502.552.602.652.702.75❂ 0 1 2 3 4 5 6 7 8initial epoch T0 [BJD +2,449,102.9199x days]-1.5-1.0-0.50.00.51.0binary period Pbin [0.086820426y days]2.452.502.552.602.652.702.75❂32503260327032803290330033103320orbital period of planet Pb [days]0.080.100.120.140.160.180.20eccentricity eb2.452.502.552.602.652.702.75❂14
K. Go´zdziewski et al.
Figure 11. Synthetic curves of the 1-planet LTT quadratic ephemeris mod-
els to optical OPTIMA measurements, including polarimetric data One of
the most deviating polarimetric points is labeled in the residuals panel.
6 RED NOISE AND/OR SYSTEMATIC ERRORS?
Analysis of the LTT observations has much in common with pul-
sar timing, planetary transits, and precision radial velocity observa-
tions, which are modelled with least-squares under the assumption
that the measurement errors are uncorrelated (white noise). How-
ever, as is known particularly by pulsar observers, the assumption
that white noise is the only source of error is unjustified when aim-
ing at estimating the underlying model parameters and their uncer-
tainties (Coles et al. 2011). In the past, this effect had been respon-
sible for false detections of planets around pulsars (Bailes et al.
1991). Similarly, correlated (red) noise or systematic errors have
been found in the planetary transit data (Pont et al. 2006) and very
recently, in the radial velocity measurements (Baluev 2011). The
same type of non-Gaussian, low-frequency correlation of residuals
to the orbital period of the binary may be present in the LTT data
collected over long time intervals.
The danger of such systematic effects in the LTT-analysed bi-
naries is reinforced due to their activity and complex astrophysical
phenomena responsible for the observed emission. One of the al-
ready well recognized mechanisms able to produce cyclic variation
of the orbital period of the binary has been proposed by Applegate
(1992). As shown by this author, a magnetic star (here, the sec-
ondary) changes its internal structure due to magnetic cycles. The
latter implies a variable zonal harmonic coefficient J2 and subse-
quently, a variable gravitational tidal field for the orbital compan-
ion which results in a varying orbital period (Hilditch 2001). The
Applegate mechanism as a possible origin of large (O-C) variations
in the HU Aqr data was studied in detail by Vogel (2008), as well
as by Schwarz et al. (2009). They discarded this possibility since
the HU Aqr stellar setup does not provide enough energy to drive
changes of the orbital period. Similar results were obtained for the
NN Ser system, that likewise has a low-mass, low-luminosity sec-
ondary star (Brinkworth et al. 2006) with a conclusion that it is
incapable of driving significant period changes in terms of the Ap-
plegate model.
Another mechanism explaining observed long-term periodic-
ities could be a slow precession of the rapidly spinning magnetic
WD star, which has been proposed as a source of long periods de-
tected in a few CVs, for instance FS Aur and V455 And (Tovmas-
sian et al. 2007). However, HU Aqr is unlikely to host such a WD,
as this AM Her-like system is known to be synchronously locked.
As a first, yet preliminary attempt, we tried to determine the char-
acteristic that can be used to quantify the shapes of the HU Aqr
light curves and might help to detect their variability and hence as-
trophysical sources of the LTT residuals. This approach mimics the
bisector velocity span (BVS) technique used to detect distortion of
spectral lines due to stellar spots and chromospheric activity. It is
well known that stellar spots may produce apparent radial velocity
changes up to 200 ms −1 (Berdyugina 2005). As a similar charac-
teristic to the BVS, we choose the slope of the linear function fitted
to the egress phase of the light curve, usually spanning no more than
a few seconds interval. We analysed 59 available light curves in the
precision OPTIMA set. The results are shown in Fig. 12. In seven
cases, we decided the data were not precise enough to derive the
slope reliably (as indicated by green filled squares) because of, for
instance, bad weather or strong wind that could introduce telescope
guidance errors. In a few other cases (seven again, marked with
blue triangles), only two points were taken for the fit, and there-
fore no error estimation was possible. Nevertheless, the obtained
slopes are uniform and span less than 2 degrees range close to 90
degrees. That furthermore indicates a similarly rapid egress phase.
The results of this test are encouraging, and support the planetary
hypothesis.
However, the slopes should be best re-computed for all avail-
able light curves that were used to determine the egress-times. The
problem of the non-homogeneity of the collected light curves still
exists. Due to varying eclipse profiles (e.g. during different accre-
tion states), the determination of mid -- egress dates is often very dif-
ficult. For instance, it could be prone to rather subjective choices
of the photometric data range to fit the parameters of the sigmoid
function, Eq. 12. That may introduce significant systematic errors,
particularly if the reduction is performed by different researchers.
This issue may be likely resolved by a re-analysis of the entire set
of all available light curves, under similar conditions paying partic-
ular attention to their origin -- the spectral window, an instrument,
and even technical and observational circumstances.
Another direction still open is a study of the binary interac-
tions, to eventually eliminate or discover astrophysical causes of the
LTT variability. The problem is in fact universal and affects other
techniques of extrasolar planets detection, such as pulsar timing
and radial velocity monitoring of active or evolved stars, as well.
It is yet possible that the observed (O-C) signal has both the plan-
etary and unmodeled astrophysical component (Potter et al. 2011),
making its unique resolution even harder.
To the best of our knowledge, possible effects of the red-noise
regarding the LTT observations have not been studied in detail.
That problem certainly deserves a deep and careful investigation.
7 CONCLUSIONS
Using a new formulation of the LTT model of the (O-C) to the avail-
able data of the HU Aqr system, we found that the 2-planet hypoth-
esis by Qian et al. (2011) is not likely. Our results reinforce recent
negative tests of dynamical stability of that system in the literature.
The self-consistent LTT model presented in this work exhibits de-
generate solutions, such as (apparently) Trojan objects of ∼ 104
Jupiter masses, or a companion in an collisional/open parabolic
or hyperbolic orbit. Ironically, two such solutions to the literature
SSQ data are the best -- fit models found in extensive, quasi-global
searches adopting a hybrid optimization.
Moreover, on the basis of a much extended, precision data set,
collected by the OPTIMA network, that increased the number of
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
-15-10 -5 0 5 10 15 20O-C [seconds]HU Aqr 1-planet LTT-fit (quadratic ephemeris, all OPTIMA), √χ2r=3.37, rms=0.79 sPb = 3205 ± 396 dayseb = 0.06 ± 0.04 -2 0 2 20 30 40 50 60 70 80residuals [s]cycle number [•103]polarimetric pointOn the HU Aquarii planetary system hypothesis
15
Figure 12. Linear slopes of 59 HU Aqr egresses derived on the basis of OPTIMA light curves by fitting a linear function only to the egress phase, usually
spanning not more than a few seconds. In seven cases, the light curves were not precise enough (as indicated by green filled squares) because of a bad weather.
In other seven cases, only two points were taken for the fit, therefore no error estimation was possible (blue triangles). See the text for more details.
data points analysed in previous works by ∼ 50 %, we have shown
that the observed (O-C) variations may be consistently explained
by the presence of only one circumbinary planet of the minimal
mass of ∼ 7 Jupiter-masses, in an orbit with a small eccentricity
of ∼ 0.1 and an orbital period of ∼ 10 years, similar to Jupiter
in the Solar system. Our results support the original 1-planet hy-
pothesis by Schwarz et al. (2009) rather than the 2-planet model
proposed by Qian et al. (2011). If confirmed, that planet would be
the next circumbinary object detected from the ground, shortly af-
ter such companions have been announced around HW Vir (Lee
et al. 2009), NN Ser (Beuermann et al. 2010), UZ For (Potter et al.
2011), SZ Her (Lee et al. 2011), DP Leo (Beuermann et al. 2011),
followed by recent discoveries of Kepler-16b (Doyle 2011), Kepler-
34b and Kepler-35b planets (Welsh et al. 2012). According to esti-
mates by these authors, the observed rate of circumbinary planets
around close binaries may be ∼ 1%.
Also, we found that the observations by Qian et al. (2011) are
not confirmed by the OPTIMA measurements due to systematic
relative shift of ∼ 3 -- 10 seconds. The nature of this discrepancy is
yet unknown. If the shift is caused by an error, all 2-planet models
presented in the literature that make use of their data are affected.
Besides the disagreement between our conclusions and the
previous works, our results suggest that the kinematic modeling of
2-planet configurations is not fully justified on the grounds of the
dynamics because the best-fit models may imply large masses (up
to stellar range), large eccentricities, and similar orbital periods in-
dicating a possibility of strong mean motion resonances. Moreover,
the (O-C) variability that suggests 2-planet solutions most likely
appears due to mixing observations done in different spectral win-
dows. That feature of the data set -- as we have shown here -- intro-
duces systematic effects that may alter the best-fit solutions signif-
icantly. This conclusion is supported by extensive numerical sim-
ulations of the 2-planet systems dynamics by Horner et al. (2011),
Wittenmyer et al. (2012) and Hinse et al. (2012). Considered within
statistical error ranges, the initial conditions lead to catastrophically
disruptive configurations, unconstrained elements of the outermost
body, and/or period damping factor β.
In this work, we found best-fit stable 2-planet models within
the quadratic ephemeris N-body model to all available data, but
the semi-major axis of the outer planet cannot be yet constrained.
Stable configurations are located within low-order MMRs spanning
tiny stable zones in the phase space, or are characterised by a large
magnitude of the period decrease. In the first case, it is difficult
to explain, how a few Jupiter mass companions could be trapped
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
in such particular, isolated resonances. In the second case, a large
β requires an efficient, internal mechanism of the binary period
change, or indicates a presence of one more companion. Our find-
ings might be a breakthrough after a few cited works reporting ba-
sically only unstable 2-planet models of the HU Aqr system, but
these discrepancies add even more ambiguity to the 2-planet hy-
pothesis.
However, the results of our experiments show that the 1-planet
solution is relatively well constrained by available optical observa-
tions selected as a homogeneous data set. Because the early opti-
cal data (the white light and V-band measurements) are coherent
with an impressive, very clear quasi-sinusoidal signal exhibited by
superior-precision OPTIMA measurements, as well as with the re-
cent MONET/N, PIRATE and WFC data, a single-companion hy-
pothesis seems well justified. A confirmation of the planetary origin
of the LTT signal still requires long-term monitoring of the system.
Due to its very long orbital period, it will take many years to con-
firm or reject the signal coherence. Such new data would be also
very useful to constrain the orbital period by the recent OPTIMA
observations alone.
Acknowledgements. We thank the anonymous referee for a
review and comments that improved the manuscript. We would like
to thank Maciej Konacki (CAMK Toru´n) for a discussion and a sug-
gestion of the egress slopes test, as well as Anna Zajczyk (CAMK,
Toru´n), Andrzej Szary (UZG, Zielona G´ora), Alex Stefanescu,
Martin Muhlegger, Helmut Steinle, Natalia Primak, Fritz Schrey,
and Christian Straubmeier (all MPE) for their help with observa-
tions. Krzysztof Go´zdziewski is supported by the Polish Ministry
of Science and Higher Education Grant No. N/N203/402739 and
POWIEW project of the European Regional Development Fund
in Innovative Economy Programme POIG.02.03.00-00-018/08. Il-
ham Nasiroglu acknowledges support from the EU FP6 Transfer
of Knowledge Project "Astrophysics of Neutron Stars" (MKTD-
CT-2006-042722). Aga Słowikowska would like to thank Bronek
Rudak for his support and discussions. She also acknowledges
support from the Foundation for Polish Science grant FNP
HOM/2009/11B, as well as from the Marie Curie European Rein-
tegration Grant within the 7th European Community Framework
Programme (PERG05-GA-2009-249168). Gottfried Kanbach ac-
knowledges support from the EU FP6 Transfer of Knowledge
Project ASTROCENTER (MTKD-CT-2006-039965) and the kind
hospitality of the Skinakas team at UoC. B. Gauza thanks the Wide
FastCam team for help in performing the observations. Research by
Tobias C. Hinse is carried out under the KRCF Young Researcher
85 86 87 88 89 90 25 30 35 40 45 50 55 60 65 70 75egress slope [degrees]cycle number [•103]HU Aqr egress slope reliablenoisy or irregular light curveonly two points used for the linear fit16
K. Go´zdziewski et al.
Fellowship Program at the Korea Astronomy and Space Science
Institute. NH acknowledges support from the NASA Astrobiology
Institute under Cooperative Agreement NNA09DA77A at the Insti-
tute for Astronomy, University of Hawaii, and from NASA/EXOB
program under grant NNX09AN05G. We thank the Skinakas Ob-
servatory for their support and allocation of telescope time. Ski-
nakas Observatory is a collaborative project of the University of
Crete, the Foundation for Research and Technology -- Hellas, and
the Max-Planck-Institute for Extraterrestrial Physics. This article
is based on observations made with the TCS telescope operated on
the island of Tenerife by the Instituto de Astrofisica de Canarias in
the Spanish Teide Observatory. This work is based in part on data
obtained with the MOnitoring NEtwork of Telescopes (MONET),
funded by the Alfried Krupp von Bohlen und Halbach Foundation,
Essen, and operated by the Georg-August-Universitat Gottingen,
the McDonald Observatory of the University of Texas at Austin,
and the South African Astronomical Observatory.
APPENDIX A: ALTERNATE MODELS TO ALL DATA
In this section, we display supplementary results illustrating a few
alternative models to the 1-planet solution of the (O-C) of the HU
Aqr binary that was analysed in the main part of the paper. Ba-
sically, all 171 data points are modeled, although in some cases,
we removed outlying data from Qian et al. (2011) because they
clearly introduce a systematic error. We considered 1-planet model
with a heuristic, sine -- damping term (Sect. A1), 2-planet kinematic
models (Sect. A2) and the full, 2-planet, self-consistent Newto-
nian model (Sect. A3). The aim of this Appendix is to demon-
strate that 2-planet models lead to non-unique or unconstrained
solutions. Hence, these results reinforce a hypothesis of a single,
quasi-sinusoidal signal of possibly planetary origin.
A1 Quadratic ephemeris 1-planet model with damping term
To describe the suggested damping signal visible in Fig. 8 (the top
right-hand panel), we modified the quadratic ephemeris model by
adding a heuristic term having the following form:
τdamp(t) = τ0 + Aexp(−t/Tdamp)sin(ndamp t + φ0),
where τ0 is an offset, A is the semi-amplitude of the signal, Tdamp is
the damping time scale, ndamp = 2π/Pdamp is the frequency, and φ0 is
the initial phase at epoch l = 0. Two examples of best-fit solutions
to all available data (171 measurements) are shown in Fig. A1. Let
us note that the planetary orbital period in the configuration in the
right-hand panel of this figure is twice the period in the 1-planet
model studied earlier. The (O-C) of a solution shown in the left-
hand panel cannot be distinguished from 2-planet models (see the
text below).
(A1)
A physical nature of that damped signal is uncertain. Allowing
for some speculations, the damping might appear due to a long --
term relaxation in the binary system which may, for instance, be
due to the binary's magnetic cycles (Applegate mechanism). In
such a case, the observed LTT signal would be resulted from two
distinct phenomena. However, we recall here that Vogel (2008) and
Schwarz et al. (2009) estimated that the Applegate mechanism can-
not be responsible for orbital changes of HU Aqr.
A2 Kinematic 2-planet models
We also tested 2-planet models with the linear and quadratic
ephemeris, (Eqs. 10, 11). Examples of the best-fit configurations
with comparable (χ2
ν)1/2 and an rms are shown in Fig. A2. Similar
to the case of the SSQ data set, no unique solution may be found.
For the parabolic ephemeris, we found many similar-quality best-
fit solutions. These fits are characterised by the orbital periods ratio
close to 1c:1b MMR with inferred planetary masses of ∼ 20 M
Jup (the bottom left-hand panel in Fig. A2), close to 4c:3b MMR
with inferred masses of ∼ 5.5 M Jup and ∼ 4.0 MJup (the top right-
hand panel in Fig. A2). A solution close to the 2c:1b MMR (the
top right-hand panel in Fig. A2), as well as configurations with ex-
treme eccentricity ec ∼ 0.95, positive damping factor β ∼ 10−12 cy-
cles day−2, and unconstrained Pc ∼ 250,000 days (not shown here)
was also found. In all these cases, the rms remains at the level of
2.4 seconds. Some of these solutions are qualitatively similar to the
2-planet fits found for the SSQ data set. These results imply that
the significantly extended data set still does not constrain 2-planet
models.
For the linear ephemeris model, we found one best -- fit solu-
tion that frequently appeared in different runs of the hybrid code.
It is shown in the bottom right-hand panel of Fig. A2. This solu-
tion is characterised by an orbital periods ratio close to 6:5. Taken
literally, this fit corresponds to Trojan brown dwarfs. However, the
kinematic model is inadequate for such a configuration of massive
objects.
A3 Newtonian, self-consistent N-body 2-planet models
In the light of the discussion presented above, we performed a pre-
liminary modelling of all available data with the help of the hybrid
algorithm driven by the self-consistent N-body model. Moreover,
we tested Lagrange stability of the best-fit models following their
orbital evolution over at least 106 orbital periods of the outermost
planet. Configurations which survived during such time without a
collision or remaining on closed orbits were regarded stable. In this
experiment we use 161 data points, excluding data in (Qian et al.
2011), due to the discrepancy with OPTIMA measurements.
To illustrate the results of the hybrid optimization, we pro-
jected the found solutions onto particular planes of the Keplerian
astrocentric, osculating elements of the planets (Fig. A3) at the
epoch of the first observation. The general finding is that the N-
body formulation helps to improve the rms, that decreased from
∼ 2.4 seconds to ∼ 1.9 seconds as compared to kinematic models.
The top row of Fig. A3 illustrates the results for the linear
ephemeris. Clearly, the data do not constrain the semi-major axes
and eccentricities of the companions. The eccentricities tend to be
large, up to 0.8. Moreover, the best-fit configuration exhibit simi-
lar values of semi-major axes (∼5.6 AU and ∼6.3 AU) and large
masses in the brown-dwarfs range of ∼ 20 Jupiter masses. We did
not find any stable configurations within this model. It is consistent
with the results for the SSQ data set (Hinse et al. 2012).
Interesting results are obtained for the quadratic ephemeris
model (see the bottom row in Fig. A2) although also this model
does not constrain orbital parameters, due to even larger spread of
the semi-major axes and eccentricities than in the linear ephemeris
ν)1/2 are found, around ab ∼ 4 au, and
model. Two minima of (χ2
ab ∼ 6 au, respectively. The best-fit configurations have (χ2
ν)1/2 ∼
2.6 and and an rms ∼ 1.9 second that is ∼ 20% better than for
ν)1/2
the best kinematic models. In the neighborhood of the first (χ2
minimum ( ab ∼ 4 AU), we found a few thousands of Lagrange
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
17
Figure A1. Examples of synthetic curves of the 1-planet LTT quadratic ephemeris models with a sine damping term to all 171 data points gathered in this
work, including 10 measurements in Qian et al. (2011), marked with white filled circles. The shaded curves represent the planetary and the damped sine signal,
respectively.
Figure A2. Examples of synthetic curves of the 2-planet LTT quadratic and linear (the bottom right panel) ephemeris models to all 171 data points analysed in
this work, including 10 data points in Qian et al. (2011) which are marked with white filled circles. The shaded curves represent single planetary signal terms,
respectively.
stable models characterised by (χ2
ν)1/2 < 3 and an rms < 2.1 (still
better than for the best 2-planet kinematic models). These fits have
well bounded ab ∼ 4 AU and small eccentricities up to 0.4. How-
ever, the osculating semi-major axis of the outer body is uncon-
strained and covers many low order mean motion resonances, be-
tween 3c:2b MMR and 5c:1b MMR. Figure A4 shows synthetic
curves of two example solutions corresponding to the 3c:2b MMR
(the left panel) and for a model close to 3c:1b MMR (the right
panel). To identify these resonances, in the neighborhoods of a
few selected best fit models, we derived high-resolution dynamical
maps (1440 ×900 data points) shown in Fig. A5. These maps are
computed in terms of the fast indicator MEGNO (Cincotta et al.
2003), with the help of our recently developed CPU cluster soft-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
-20-10 0 10 20O-C [seconds]HU Aqr 1-planet LTT-fit (quad. ephem. + damped sine, all data), √χ2r=2.6, rms=2.4 sPb = 3418 ± 28 dayseb = 0.16 ± 0.03 Pdamp = 2048 days -8 0 8 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-40-30-20-10 0 10 20 30 40 50O-C [seconds]HU Aqr 1-planet LTT-fit (quad. ephem. + damped sine, all data), √χ2r=3.6, rms=2.6 sPb = 6656 ± 22 dayseb = 0.00 ± 0.02 Pdamp = 2824 days-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-15-10 -5 0 5 10 15 20 25O-C [seconds]HU Aqr 2-planet LTT-fit (quadratic ephemeris, all data), √χ2r=2.41, rms=2.35 sPb = 3048 ± 54 dayseb = 0.19 ± 0.07Pc = 4012 ± 356 daysec = 0.60 ± 0.05-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-20-10 0 10 20 30 40 50O-C [seconds]HU Aqr 2-planet LTT-fit (quadratic ephemeris, all data), √χ2r=2.42, rms=2.35 sPb = 3053 ± 33 dayseb = 0.25 ± 0.01Pc = 7788 ± 165 daysec = 0.68 ± 0.05-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-60-40-20 0 20 40 60O-C [seconds]HU Aqr 2-planet LTT-fit (quadratic ephemeris, all data), √χ2r=2.50, rms=2.36 sPb = 4382 ± 42 dayseb = 0.11 ± 0.02Pc = 4796 ± 50 daysec = 0.15 ± 0.02-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-80-60-40-20 0 20 40 60O-C [seconds]HU Aqr 2-planet LTT-fit (linear ephemeris, all data), √χ2r=2.53, rms=2.36 sPb = 5132 ± 35 dayseb = 0.09 ± 0.01Pc = 6172 ± 89 daysec = 0.32 ± 0.01-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]18
K. Go´zdziewski et al.
Figure A6. Statistics of 2-planet N-body quadratic ephemeris models gath-
ered with the hybrid algorithm, projected onto the (ac,β) -- plane, for the
quadratic ephemeris. Meaning of symbols is the same as in Fig. A3.
ware MECHANIC3 (Słonina et al. 2012). MEGNO measures the
maximal Lyapunov exponent, which makes it possible to distin-
guish between chaotic and regular solutions. Each point at these
maps has been integrated over ∼ 104 orbital periods of the outer-
most companion. The dynamical maps confirm that the Lagrange
stable models examined over a limited time-span are equivalent to
quasi-periodic, stable solutions.
A solution illustrated in the left panel of Fig. A4 is the best-
ν)1/2 ∼ 2.82
fit stable model found in the hybrid search with (χ2
and an rms∼ 2 sec. It is located in a very narrow, isolated stabil-
ity island of the 3c:2b MMR and characterised by relatively large
β ∼ −2.7× 10−13 day cycle−2, similarly to the kinematic model.
The right panel of Fig. A4 shows a configuration close to the
3c:1b MMR, which has even larger β ∼ −6 × 10−13 day cycle−2.
Figure A6 shows a statistics of the best-fit solutions in the
(ac,β) -- plane. It reveals that β is not constrained, regarding even
its sign. Stable models exhibit a strong correlation between both
these parameters. A larger value of the semi-major axis of the outer
planet is related to a larger magnitude of β. Due to this correlation,
an interpretation of stable configurations is complex. For relatively
small magnitude of β, stable configurations are characterized by
low order MMRs and may be found in tiny areas of stable motions
(see Fig. A5). For more separated planets, when stability zones
are much more extended, β increases. Already β ∼ 2 × 10−13
day cycle−2 is difficult to explain by physical phenomena in the
binary, as we discussed in Sect. 5. Such large values of β may in-
dicate a third, long-period companion object in a very distant orbit.
However, because already the 2-planet model is not constrained by
the data, also a 3-planet configuration cannot be fixed without am-
biguity. We did an attempt to search for such Newtonian 3-planet
models within the linear ephemeris, but we did not find any im-
proved, nor stable solutions of this type.
REFERENCES
Applegate J. H., 1992, ApJ, 385, 621
Bailes M., Lyne A. G., Shemar S. L., 1991, Nature, 352, 311
Baluev R. V., 2011, Celestial Mechanics and Dynamical Astron-
omy, 111, 235
3 http://git.astri.umk.pl/projects/mechanic
Beard S. M., Vick A. J. A., Atkinson D., Dhillon V. S., Marsh T.,
McLay S., Stevenson M., Tierney C., 2002 Vol. 4848 of Society
of Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, Ultracam camera control and data acquisition system. pp
218 -- 229
Berdyugina S. V., 2005, Living Reviews in Solar Physics, 2, 8
Beuermann K., Buhlmann J., Diese J., Dreizler S., Hessman
F. V., Husser T.-O., Miller G. F., Nickol N., Pons R., Ruhr D.,
Schmulling H., Schwope A. D., Sorge T., Ulrichs L., Winget
D. E., Winget K. I., 2011, A&A, 526, A53
Beuermann K., Hessman F. V., Dreizler S., Marsh T. R., Parsons
S. G., Winget D. E., Miller G. F., Schreiber M. R., Kley W.,
Dhillon V. S., Littlefair S. P., Copperwheat C. M., Hermes J. J.,
2010, A&A, 521, L60
Brinkworth C. S., Marsh T. R., Dhillon V. S., Knigge C., 2006,
MNRAS, 365, 287
Charbonneau P., 1995, A&A Suppl., 101, 309
Cincotta P. M., Giordano C. M., Sim´o C., 2003, Physica D Non-
linear Phenomena, 182, 151
Coles W., Hobbs G., Champion D. J., Manchester R. N., Verbiest
J. P. W., 2011, MNRAS, 418, 561
Dhillon V. S., et al., 2007, MNRAS, 378, 825
Doyle L. R. e. a., 2011, Science, 333, 1602
Eastman J., Siverd R., Gaudi B. S., 2010, PASP, 122, 935
Go´zdziewski K., Konacki M., 2004, ApJ, 610, 1093
Go´zdziewski K., Konacki M., Maciejewski A. J., 2003, ApJ, 594,
1019
Go´zdziewski K., Maciejewski A. J., 2001, ApJL, 563, L81
Go´zdziewski K., Migaszewski C., Musieli´nski A., 2008, in Y.-
S. Sun, S. Ferraz-Mello, & J.-L. Zhou ed., IAU Symposium
Vol. 249 of IAU Symposium, Stability constraints in modeling
of multi-planet extrasolar systems. pp 447 -- 460
Hairer E., Naersett S. P., Wanner G., 2009, Solving Ordinary Dif-
ferential Equations I: Nonstiff Problems
Hellier C., 2001, Cataclysmic Variable Stars
Hilditch R. W., 2001, An Introduction to Close Binary Stars
Hinse T. C., Lee J. W., Go´zdziewski K., Haghighipour N., Lee
C.-U., Scullion E. M., 2012, MNRAS, p. 2221
Holmes S., Kolb U., Haswell C. A., Burwitz V., Lucas R. J., Ro-
driguez J., Rolfe S. M., Rostron J., Barker J., 2011, PASP, 123,
1177
Horner J., Marshall J. P., Wittenmyer R. A., Tinney C. G., 2011,
MNRAS, 416, L11
Irwin J. B., 1952, ApJ, 116, 211
Kanbach G., Kellner S., Schrey F. Z., Steinle H., Straubmeier C.,
Spruit H. C., 2003 Vol. 4841 of Society of Photo-Optical In-
strumentation Engineers (SPIE) Conference Series, Design and
results of the fast timing photo-polarimeter OPTIMA. pp 82 -- 93
Kanbach G., Stefanescu A., Duscha S., Muhlegger M., Schrey F.,
Steinle H., Slowikowska A., Spruit H., 2008 Vol. 351 of Astro-
physics and Space Science Library, OPTIMA: A High Time Res-
olution Optical Photo-Polarimeter. pp 153 -- +
Laughlin G., Chambers J. E., 2001, ApJL, 551, L109
Lee J. W., Kim S.-L., Kim C.-H., Koch R. H., Lee C.-U., Kim
H.-I., Park J.-H., 2009, AJ, 137, 3181
Lee J. W., Lee C.-U., Kim S.-L., Kim H.-I., Park J.-H., 2011,
ArXiv e-prints
Lee M. H., Peale S. J., 2003, ApJ, 592, 1201
Malhotra R., 1993, in J. A. Phillips, S. E. Thorsett, & S. R. Kulka-
rni ed., Planets Around Pulsars Vol. 36 of Astronomical So-
ciety of the Pacific Conference Series, Orbital dynamics of
PSR1257+12 and its two planetary companions. pp 89 -- 106
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
On the HU Aquarii planetary system hypothesis
19
Figure A3. Statistics of 2-planet N-body models gathered with the hybrid algorithm, projected onto planes of selected parameters. The top row illustrates
the results for the linear ephemeris, the bottom row shows the fits for the quadratic ephemeris model. The rms quality of these solutions is coded as filled
circles: the better fit -- the darker colour. Orbital parameters of the best-fit solutions are marked with shaded, intersecting lines and a flower symbol. Solutions
Lagrange -- stable over 106 revolutions of the outermost planet are marked with red -- white circles.
Figure A4. Examples of synthetic LTT curves and (O-C) residuals for stable 2-planet quadratic ephemeris Newtonian (N -- body) models to 161 data points
analysed in this work, without 10 data points in Qian et al. (2011). These models are selected from a sample illustrated in Fig. A3. Dynamical maps of these
solutions shows Fig. A5 (they are labelled with the Roman numerals, as I and IV, respectively).
Morbidelli A., 2002, Modern celestial mechanics : aspects of solar
system dynamics
Nasiroglu I., Słowikowska A., Kanbach G., Schwarz R., Schwope
A. D., 2010, in High Time Resolution Astrophysics (HTRA) IV -
The Era of Extremely Large Telescopes Proceedings of Science,
The orbital ephemeris of HU Aquarii observed with OPTIMA.
Are there two giant planets in orbit?
Pont F., Zucker S., Queloz D., 2006, MNRAS, 373, 231
Potter S. B., et al., 2011, MNRAS, 416, 2202
Press W. H., 2002, Numerical recipes in C++ : the art of scientific
computing
Qian S.-B., Liu L., Liao W.-P., Li L.-J., Zhu L.-Y., Dai Z.-B., He
J.-J., Zhao E.-G., Zhang J., Li K., 2011, MNRAS, 414, L16
Schwarz R., Schwope A. D., Vogel J., Dhillon V. S., Marsh T. R.,
Copperwheat C., Littlefair S. P., Kanbach G., 2009, A&A, 496,
833
Schwope A., et al., 2004, in S. Vrielmann & M. Cropper ed., IAU
Colloq. 190: Magnetic Cataclysmic Variables Vol. 315 of As-
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
2 4 6 8 10 12 14 16 18 20 2 4 6 8 10 12 14 16semi-major axis ac [au] semi-major axis ab [au]rms<2.1srms<2.0srms<1.9s❁ 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1eccentricity ececcentricity ebrms<2.1srms<2.0srms<1.9s❁-20-15-10 -5 0 5 10 15 20O-C [seconds]HU Aqr 2-planet LTT-fit (N-body, quad. eph., ALL data), √χ2r=2.82, rms=2.0 sab = 3.71 aueb = 0.05 ac = 4.75 auec = 0.23 β = -2.7•10-13 day cl-2I-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]-50-40-30-20-10 0 10 20 30 40 50O-C [seconds]HU Aqr 2-planet LTT-fit (N-body, quad. eph., ALL data), √χ2r=2.97, rms=2.1 sab = 3.99 aueb = 0.07 ac = 8.24 auec = 0.03 β = -6.3•10-13 day cl-2IV-10 0 10 0 10 20 30 40 50 60 70 80residuals [s]cycle number [•103]20
K. Go´zdziewski et al.
Figure A5. MEGNO dynamical maps in the (ac,ec) -- plane for a few representative N-body stable solutions illustrated in the bottom panels of Fig. A4. Yellow
colour encodes strongly unstable (chaotic) configurations, and purple colour (MEGNO (cid:104)Y(cid:105) ∼ 2) is for stable, quasi-periodic solutions. Parameters of the
nominal, tested fits are marked with the star symbol. The most prominent, low-order mean motion resonances are labeled. The original resolution of these
dynamical maps is 1440× 900 data points integrated for 104 outermost orbital period each. The total mass of the binary is 0.98 M(cid:12) (Schwope et al. 2011).
tronomical Society of the Pacific Conference Series, REVIEW:
Multiwavelength Observations of eclipsing polars. p. 92
Schwope A. D., Horne K., Steeghs D., Still M., 2011, A&A, 531,
A34
Schwope A. D., Schwarz R., Sirk M., Howell S. B., 2001, A&A,
375, 419
Słonina M., Go´zdziewski K., Migaszewski C., 2012, in F. Are-
nou & D. Hestroffer ed., Orbital Couples: Pas de Deux in the
Solar System and the Milky Way (arXiv:1202.6513v1,in print)
Mechanic: a new numerical MPI framework for the dynamical
astronomy
Tovmassian G. H., Zharikov S. V., Neustroev V. V., 2007, ApJ,
655, 466
Vogel J., 2008, PhD thesis, Technischen Universitat Berlin, Ger-
many
Vogel J., Schwope A., Schwarz R., Kanbach G., Dhillon V. S.,
Marsh T. R., 2008, in D. Phelan, O. Ryan, & A. Shearer ed., High
Time Resolution Astrophysics: The Universe at Sub-Second
Timescales Vol. 984 of American Institute of Physics Confer-
ence Series, On the orbital period of the magnetic cataclysmic
variable HU Aquarii. pp 264 -- 267
Warner B., 1995, Cambridge Astrophysics Series, 28
Welsh et al. 2012, Nature, doi:10.1038/nature10768 (in press), 1
Wittenmyer R. A., Horner J., Marshall J. P., Butters O. W., Tinney
C. G., 2012, MNRAS, 419, 3258
c(cid:13) 0000 RAS, MNRAS 000, 000 -- 000
|
1205.0822 | 2 | 1205 | 2012-09-13T23:30:12 | A dynamical analysis of the Kepler-11 planetary system | [
"astro-ph.EP"
] | The Kepler-11 star hosts at least six transiting super-Earth planets detected through the precise photometric observations of the Kepler mission (Lissauer et al.). In this paper, we re-analyze the available Kepler data, using the direct N-body approach rather than an indirect TTV method in the discovery paper. The orbital modeling in the realm of the direct approach relies on the whole data set, not only on the mid-transits times. Most of the results in the original paper are confirmed and extended. We constrained the mass of the outermost planet g to less than 30 Earth masses. The mutual inclinations between orbits b and c as well as between orbits d and e are determined with a good precision, in the range of [1,5] degrees. Having several solutions to four qualitative orbital models of the Kepler-11 system, we analyze its global dynamics with the help of dynamical maps. They reveal a sophisticated structure of the phase space, with narrow regions of regular motion. The dynamics are governed by a dense net of three- and four-body mean motion resonances, forming the Arnold web. Overlapping of these resonances is a main source of instability. We found that the Kepler-11 system may be long-term stable only in particular multiple resonant configurations with small relative inclinations. The mass-radius data derived for all companions reveal a clear anti-correlation between the mean density of the planets with their distance from the star. This may reflect the formation and early evolution history of the system. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- 21 (2012)
Printed 4 November 2018
(MN LATEX style file v2.2)
A dynamical analysis of the Kepler-11 planetary system
Cezary Migaszewski1(cid:63), Mariusz Słonina1† and Krzysztof Go´zdziewski1(cid:63)‡
1Toru´n Centre for Astronomy, Nicolaus Copernicus University, Gagarin Str. 11, 87-100 Toru´n, Poland
Accepted 2012 August 23. Received 2012 August 20; in original form 2012 May 3
2
1
0
2
p
e
S
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
2
2
8
0
.
5
0
2
1
:
v
i
X
r
a
ABSTRACT
The Kepler-11 star hosts at least six transiting super-Earth planets detected through the pre-
cise photometric observations of the Kepler mission (Lissauer et al.). In this paper, we re-
analyze the available Kepler data, using the direct N-body approach rather than an indirect
TTV method in the discovery paper. The orbital modeling in the realm of the direct approach
relies on the whole data set, not only on the mid -- transits times. Most of the results in the orig-
inal paper are confirmed and extended. We constrained the mass of the outermost planet g to
less than 30 Earth masses. The mutual inclinations between orbits b and c as well as between
orbits d and e are determined with a good precision, in the range of [1,5] degrees. Having sev-
eral solutions to four qualitative orbital models of the Kepler-11 system, we analyze its global
dynamics with the help of dynamical maps. They reveal a sophisticated structure of the phase
space, with narrow regions of regular motion. The dynamics are governed by a dense net of
three -- and four -- body mean motion resonances, forming the Arnold web. Overlapping of these
resonances is a main source of instability. We found that the Kepler-11 system may be long-
term stable only in particular multiple resonant configurations with small relative inclinations.
The mass-radius data derived for all companions reveal a clear anti-correlation between the
mean density of the planets with their distance from the star. This may reflect the formation
and early evolution history of the system.
Key words: celestial mechanics -- planetary transits -- Kepler-11 -- Arnold web
1
INTRODUCTION
The Kepler space mission is a breakthrough in the field of searches
for the Earth -- like extrasolar planets (Borucki et al. 2010; Koch
et al. 2010; Jenkins et al. 2010; Caldwell et al. 2010). About of
150,000 solar dwarfs are monitored by 0.95 -- meter Kepler tele-
scope. The photometric data are publicly available from the MAST
archive1.
To date, the mission identified more that 2,200 planetary can-
didates (Batalha et al. 2012). Among them, many multi-planet sys-
tems are found. For instance, planets were confirmed in two -- planet
configurations, i.e., Kepler-10 (Batalha et al. 2011; Fressin et al.
2011), Kepler-25, 26, 27, 28 (Steffen et al. 2012), Kepler-29, 31, 32
(Fabrycky et al. 2012), Kepler-23, 24 (Ford et al. 2012); in three-
planet systems Kepler-9 (Holman et al. 2010), Kepler-30 (Fabrycky
et al. 2012), Kepler-18 (Cochran et al. 2011); in four -- planet sys-
tems (Borucki et al. 2011), as well as in five -- planet configurations
Kepler-20 (Gautier et al. 2011; Fressin et al. 2011), Kepler-33 (Lis-
sauer et al. 2012). The Kepler-11 hosts six planetary companions
(Lissauer et al. 2011). The transiting planet candidates can be con-
firmed through determining their masses with the help of the so-
(cid:63) E-mail: [email protected]
† E-mail: [email protected]
‡ E-mail: [email protected]
1 http://archive.stsci.edu/kepler
c(cid:13) 2012 RAS
called Transit Timing Variations method (TTV, Holman & Mur-
ray 2005; Agol et al. 2005). In this approach, the (O-C) variations
between observed mid-transit times and their ephemeris are the
observables, which can be fitted by an appropriate orbital model.
In recent papers, also additional observables are analysed, like the
so-called Transits Duration Variations (TDVs) (see, e.g., Nesvorn´y
et al. 2012).
In this paper, we re-analyse the photometric data of Kepler-11
with a modified, direct approach providing an alternate estimation
of masses and orbital elements. To describe this method further
in the paper, we recall shortly the main conclusions in (Lissauer
et al. 2011). Using the TTV method and an assumption of strictly
coplanar model of the system, they determined masses of five inner
planets in the range of a few Earth masses. The outermost planet
interacts weakly with the inner companions, and its mass could be
roughly constrained as smaller than the Jupiter mass. It has been
not confirmed as a planet, although the probability of blending is
very small, ∼ 0.001. Orbital eccentricities in the Kepler-11 system
were determined only for the five inner objects. Due to the assump-
tion of coplanarity, a determination of mutual inclinations between
the orbits was not possible. Lissauer et al. (2011) argue that these
inclinations should remain in the range of [0,2] degrees. The dy-
namical analysis have revealed that the system is not involved in
the mean motion resonances (MMRs), however a pair of planet b
and planet c is close to 5:4 MMR.
The determined masses and radii of the planets imply con-
2
C. Migaszewski, M. Słonina and K. Go´zdziewski
strains on their chemical composition. Planets d, e and f might have
similar internal compositions to those of Uranus or Neptune, while
planets b and c are rather ice rich, with a smaller amount of H2/He
mixture than these planets in the Solar system.
In this paper, we focus mostly on the global dynamics of the
system and a few aspects which were not addressed in the discovery
paper.
First of all, we model the available Kepler data through a di-
rect algorithm that relies on the self-consistent N-body fitting of the
light-curves, instead of the TTV method applied in the discovery
work. The TTV algorithm makes use of the transit times a poste-
riori, after they are determined from the light curves. Through ex-
tensive numerical experiments, we found that the direct approach
brings more information than the TTV method. For instance, we
could constrain the mass of the outermost planet to less than ∼
30 Earth masses. We also found significant bounds for the mutual
inclinations to less than 5◦ for planets b and c as well as for plan-
ets d and e.
The direct model, also called the dynamical-photometric
model, already was used in a few papers. For instance, it was ap-
plied to analyse the light curve of the triple-star system KOI-126
(Carter et al. 2011), and to estimate masses of two planets transit-
ing Kepler-36 (Carter et al. 2012). This algorithm also verified the
Kepler-9 model, which was found first with the help of the TTV
algorithm (Holman et al. 2010).
A number of initial conditions found with the direct approach
makes it possible to investigate the dynamics of the system. We
focus on the short -- term time scale, governed by the mean mo-
tion resonances. We study the multi-dimensional structure of the
phase space with the help of dynamical maps. In the vicinity a few
qualitative transit models considered in this work, the dynamics are
governed by a dense net of 3 -- body and 4 -- body mean motion reso-
nances. This net may be identified with the Arnold web, which is a
feature of close to integrable Hamiltonian systems. The Kepler-11
appears as strongly resonant extrasolar system, and this feature may
reflect its trapping into MMRs at the early stages of the formation
and evolution.
Using a new determination of the masses and radii, we found
a curious mass-radius relation implying a clear anti-correlation be-
tween the mean density of the planets and their distances from
the star. Their densities exhibit a sequence of planet b which is
denser than Neptune, through the Neptune-like planet c, Uranus-
like planet d, Jupiter-like planets e and f, and planet g which is
likely Jupiter/Saturn-like.
The paper is structured as follows. In Sect. 2, we shortly de-
scribe the photometric data of Kepler-11 available in the MAST
archive. We also refine the observational TTV model. In Sect. 3, we
present the results derived through intensive computations with the
bootstrap algorithm. Furthermore, we discuss a possible composi-
tion of the planets (Sect. 4). Section 5 is devoted to the dynamical
analysis of the Kepler-11 system. Conclusions and prospects for a
future work are given at the end of this paper.
2 TRANSITS IN A MULTI-PLANET SYSTEM
The photometric data of Kepler-11 were taken from the MAST
archive. At the time of writing this paper, the publicly available
light-curves span about of 500 days in six parts. These data were
binned on ∼ 30-minute intervals. We analysed a "de-trended" data
set derived through a smoothing procedure. At first, we isolate all
transits from the light curve. Then the moving average with a time-
step of 0.5 days provides the mean level of the flux. Next, we con-
struct an interpolated, reference light curve with the cubic spline on
these nodes. Finally, we divide the raw flux, with all transits data,
by its values of the reference, mean level flux curve.
The de-trended data available in the MAST database exhibit
a growth of the flux shortly before and after a particular transit. In
some parts of the available light-curves, spanning approximately
300 days, the measurements appear in the raw form. We did not
use these data, aiming to analyze a possibly uniform set of obser-
vations.
2.1 Modeling the stellar flux
A common model of photometric observations of a star transited
by planetary companions consists of two major parts. The first part
concerns the flux deficit due to small, dark objects passing in front
of the star. At first, the average orbital periods are determined. Then
transit depths and duration times are parametrized on the basis of
phase -- folded light-curves. Single mid-transit times are also deter-
mined. At the second level, we can estimate the planetary masses
and orbital elements fitting a model of motion of mutually interact-
ing planets.
We focus on the first level of the photometric analysis. To
compute the flux deficit, we use the quadratic limb darkening model
(Mandel & Agol 2002), recalling that the Kepler-11 light-curves
are relatively noisy and sampled with a low frequency,
∆I(r) = 1− γ1 (1− cosθ)− γ2 (1− cosθ)2 ,
(1)
where r is the normalized radial coordinate w.r.t. the centre of the
stellar disk, θ is the angle between the direction to the observer and
the normal to the stellar surface. The two limb-darkening coeffi-
cients γ1 and γ2 must be positive and γ1 +γ2 < 1 (see a study of the
limb darkening coefficients for a few target stars of the Kepler mis-
sion, Howarth 2011). For small ratio p ≡ Rp/Rs of planet radius
Rp to the stellar radius Rs, Mandel & Agol (2002) found an ana-
lytic approximation of the flux deficit, ∆F = ∆F(z; p,γ1,γ2) which
depends on the normalized distance z between the centers of stellar
and planetary disks, projected onto the sky plane (see Eq. 8 in the
cited paper), as well as on p and γ1,2.
If more than one planet transits the star at the same time, the
total flux deficit can be computed as the sum of the deficits caused
by particular planets. Obviously, γ1,γ2 are the same for all plan-
ets, while p and z are different for each object. If transiting plan-
ets are small, we can use a simple model of independent transits
rather than more general treatment (e.g., P´al 2011). Because we
model the photometric measurements directly, by reconstructing
the whole light-curve, we are not restricted to single transits and
mid-transit times. Also multiple transits can be covered. In light of
relatively narrow observational window, multiple transits are very
helpful to constrain orbital elements of the transit model.
Figure 1 displays a few selected fragments of the data set
marked with red dots and error bars which are over-plotted on the
synthetic curve best -- fitting the data (blue curve). The fitting pro-
cedure will be described in more detail in Sect. 2.3. The last panel
shows transits of three planets (b, d and e).
2.2 The model of orbital motion
The orbital motion of multiple planetary system is described in
terms of the full N-body problem in the Poincar´e reference frame
(e.g., Morbidelli 2002). In this frame, the Cartesian coordinates
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
3
Figure 1. Sample synthetic light curves over-plotted on photometric measurements of Kepler-11. Red dots are for the observational data, blue solid curve is
for synthetic light-curve derived in this work. The reference epoch is JD 2,455,500.
of the planets are astrocentric, while their velocities are barycen-
tric. The equations of motion are integrated with the second order
symplectic integrator SABA2 (Wisdom & Holman 1991; Laskar
& Robutel 2001). It provides 2-3 times better CPU performance
than other algorithms, which we tested (like the Bulirsh-Stoer-
Gragg scheme, BGS) constrained with the same time -- step accu-
racy. To speed-up the computations even more, we did not integrate
the system at all measurements moments. This would force ∼ 30-
minute step-size of the integrator. Instead, we fixed this step-size to
∆t ∼ 1/20 of the innermost period of planet b, i.e., ∆t ≈ 0.5 day.
Furthermore, the flux function F(t) is computed only close to
the mid-transits. Ingress and egress times of particular events are
tabulated. When a transit takes place, the coordinates of particu-
lar planet at time t required to evaluate the flux deficit are deter-
mined through the polynomial interpolation on five nodes around
t. Through a comparison with the direct, full-accuracy integrations
with the BGS algorithm, we found that the selected time time-step
and the number of interpolation nodes provide a sufficient preci-
sion and acceptable CPU overhead. We examined this method by
changing the number of nodes in the polynomial interpolation, as
well as the time step-size. The flux level, interpolated on five nodes
and with ∆ t ≈ 0.5 day, differs from its exact value by less than
10−9.
2.3 Optimization algorithm and error estimation
We searched for the best -- fit model of the transits by a common
minimization of the χ2
ν function. This function is defined as fol-
lows:
(cid:104)
(cid:105)2
χ2
ν =
1
Nobs − Np − 1
Nobs∑
j=1
1
σ2
j
Fj − F(t j)
,
(2)
where Nobs is the number of observations, Np is the number of free
parameters, ν = Nobs − Np − 1 is the number of the degrees of free-
dom, σ j is the error of the j-th observation Fj, and F(t j) is a model
function evaluated at time t j. This form of the χ2
ν -- function is cor-
rect if the uncertainties are uncorrelated (see, e.g., Baluev 2009).
To verify whether the available photometric data fulfill this assump-
tion, one has to use a more general statistical model incorporat-
ing the red -- noise effect. However, under particular settings of our
N-body photometric model, this would require an enormous CPU
overhead. Hence, we use equation 2 as a reasonable first order ap-
proximation.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
The best -- fit parameters of the transits model are searched
through a two -- step optimization method. In the first step, we ap-
ply a robust and well tested quasi-global Genetic Algorithm (GA,
see, Charbonneau 1995; Deb 2004)2 which makes it possible to
find promising solutions. A local, fast gradient method (here, the
Levenberg-Marquardt algorithm) is then used to refine the solu-
tions found in the GA step. Such an approach is called the hybrid
optimization (see Go´zdziewski et al. 2008, and references therein).
Let us note that the parameter space is huge as it has dimension
of 50. Some of these parameters can be determined very well, like
the orbital periods of the transiting planets. Unfortunately, due to
the relatively short observational time window, many parameters
which are critical for the stability (relative inclinations, masses,
nodal lines) cannot be well constrained. It makes the fitting pro-
cess a challenging problem.
The parameter errors are estimated through the bootstrap al-
gorithm (see, e.g., Press et al. 1992). The bootstrap is CPU --
demanding, but it is straightforward method to estimate standard
errors in high-dimensional problems and for large number of data.
The light -- curves which we analyzed have ∼22,000 points. The
bootstrap algorithm requires to find the best -- fit solutions to a large
number of synthetic sets derived through random sampling with re-
placement from the original measurements. To obtain reliable error
estimates of the best -- fit parameters, one needs at least ∼ 103 -- 104
synthetic solutions. When such a large set of the best -- fit models
is gathered, we constructed normalized histograms for each free
parameter. These histograms reflect the parameter distribution in
response to the errors of the measurements, and may be smoothly
approximated by an asymmetric Gaussian function. This makes it
possible to determine the standard uncertainties. To perform the
bootstrap procedure, at first one needs to find reliable best -- fit pa-
rameters for the nominal data set. This step was done through an
intensive quasi-global search with the help of the hybrid algorithm.
The bootstrap computations are CPU-time consuming and were
performed on the reef CPU-cluster of the Pozna´n Supercomput-
ing Centre.
2 We use publicly available implementation by Kalyanmoy Deb, see http:
//www.iitk.ac.in/kangal/pub.htm
cd0.9980.9991.0001.001 548.9 549.1 549.3 549.5 549.7t[d]fluxfb0.9980.9991.0001.001 697.6 697.8 698 698.2 698.4t[d]fluxdb0.9980.9991.0001.001 707.9 708.1 708.3 708.5 708.7t[d]fluxcf0.9980.9991.0001.001 744.2 744.4 744.6 744.8 745t[d]fluxdb0.9980.9991.0001.001 821.3 821.5 821.7 821.9 822.1t[d]fluxbde0.9970.9980.9991.0001.001 934.7 934.9 935.1 935.3 935.5t[d]flux4
C. Migaszewski, M. Słonina and K. Go´zdziewski
2.4 Numerical setup of the dynamical analysis
In spite of small eccentricities and apparently co-planar orbits, the
Kepler-11 systems is orbitally very active. It appears as dynam-
ically packed planetary system (the definition is given in Barnes
et al. 2008), with only narrow stable zones in the phase space. For
this reason, we used the best-fit model solutions gathered in the
bootstrap search as the input data to extensive dynamical study of
this system. As we will discuss later, a study of the stability is a
challenging problem. Due to relatively short observational window
(∼ 500 days), weak transits having depths comparable with the
measurements errors and a small number of data points covering
particular transits (typically 10− 15), the derived initial conditions
may be shifted away from the real configurations.
To investigate the dynamics of the Kepler-11 system in a
global manner, we applied an approach in our previous papers
which is well established in the literature. It relies on reconstruct-
ing the structure of the phase space with the fast indicator MEGNO
(Cincotta & Sim´o 2000; Cincotta et al. 2003). This dynamical char-
acteristic makes it possible to distinguish between regular (stable)
and irregular (chaotic, unstable) trajectories in the phase space by
computing relatively short numerical orbits. Having representative
solutions selected in the bootstrap statistics, we study their neigh-
borhood on the dynamical maps. Constructing a dynamical map
relies on two model parameters, e.g., the semi-major axes of a pair
of planets. The selected parameters are varied in the given range at
a discrete grid. The remaining components of the initial parameter
vector are fixed at their nominal values. If it is necessary, they are
altered to preserve the observational constraints. Then we calculate
MEGNO at each point of the grid. Dynamical maps are informa-
tive and become a standard numerical tool helpful to understand
the global dynamics of multiple systems.
To compute the MEGNO indicator, we must solve the varia-
tional equations to the equations of motion of the planetary N-body
problem. The Kepler-11 system architecture with low -- eccentric or-
bit and small masses is an ideal target for an efficient symplectic al-
gorithm described in (Go´zdziewski 2003; Go´zdziewski et al. 2008).
The general-purpose integrators, like the Runge-Kutta or Bulirsh-
Stoer-Gragg schemes are not efficient nor accurate enough in this
case. These methods introduce a systematic drift of the energy and
other integrals. To avoid such errors, and to solve the variational
equations, we apply the tangent map introduced by Mikkola & In-
nanen (1999). As the very basic step, it requires to differentiate
the "drift" and "kick" maps of the standard leap -- frog algorithm.
The variations may be then propagated within the same symplectic
scheme, as the equations of motion. Having the variational vector
δδδ computed at discrete times, we find temporal y and mean Y of
the MEGNO at the j-th integrator step j = 1,2, . . ., (Cincotta et al.
2003; Go´zdziewski et al. 2008):
Y ( j) =
( j− 1)Y ( j− 1) + y( j)
j
,
(cid:18) δ j
(cid:19)
j− 1
j
y( j− 1) + 2ln
y( j) =
with initial conditions y(0) = 0, Y (0) = 0, δ = δδδ. The MEGNO
maps tend asymptotically to
δ j−1
Y ( j) = ah j + b,
where a = 0,b ∼ 2 for quasi-periodic orbits, a = b = 0 for stable,
periodic orbit, and a = (1/2)σ,b = 0 for chaotic orbit with the max-
imal Lyapunov exponent σ. The tangent MEGNO map is linear,
hence the variational vector can be normalized, if its value grows
too large for chaotic orbits. In practice, we stop the integration if
the MEGNO indicator reaches a given limit (usually, Y = 5).
The symplectic maps were propagated with the 4-th order
SABA4 scheme in (Laskar & Robutel 2001). A choice of the
fixed step-size must be carefully controlled. We did this, checking
whether the relative energy error is "flat" across the dynamical map
(Go´zdziewski et al. 2008) and sufficiently small. Indeed, the step --
size ∼ 0.5 day preserved this error at a level of 10−11 over the total
integration times up to T ∼ 40,0000 yr (∼ 100,000 periods of the
outermost planet). This time scale is long enough to detect the most
significant 2-body and 3 -- body MMRs though even such integration
period may be insufficient to detect all "dangerous" unstable reso-
nances. Weakly chaotic motions due to multi-body MMRs still may
lead to catastrophic events after much longer time (Go´zdziewski
et al. 2008).
The dynamical maps in this paper have typical resolution up
to 512× 512 pixels. This requires an enormous CPU-time. It is ba-
sically not possible to perform such intensive computations on a
single workstation. Therefore, we used our new Message Passing
Interface (MPI) based environment Mechanic (Slonina et al. 2012)
to perform the computations in a reasonable time in CPU-clusters3.
They were performed on the reef cluster at the Pozna´n Supercom-
puting Centre. A Mechanic run of a typical dynamical map occu-
pied up to 1200 CPU cores for ∼ 16 hours.
2.5 Free parameters of the transit model
The free parameters of the transit model are the stellar radius
R0, the limb darkening coefficients γ1,γ2; the mass mi, radius
Ri and orbital elements of each planet
in the system, where
i =b,c,d,e,f,g. Planetary orbits are described through the Poincar´e
geometric, osculating elements at the epoch of the first observation
JD 2455964.51128: a tuple (ai,ei,Ii,Ωi,ωi,Mi) is for the semi-
major axis, eccentricity, inclination to the plane of the sky, the
longitude of ascending nodes, the argument of pericenter, and the
mean anomaly, respectively. The orbital node of the first planet,
Ωb = 0◦ due to invariance of the model with respect to a rotation
of the whole system. The inclinations are obviously close to 90◦. A
deviation from 90◦ is irrelevant only for single-planet systems.
In a multi-planet system, some orbits may be inclined to the
sky plane by angle (cid:54)= 90◦, which implies different relative incli-
nations between orbits of particular planets, even for the same
longitudes of nodes. Due to the invariance of transits with re-
spect to the direction of the total angular momentum of the sys-
tem, a combination of (Ii (cid:54) 90◦,Ωi) means the same geometry as
([180◦−Ii] (cid:62) 90◦,−Ωi). Thus, when necessary, for a given planet p
we can fix the range of Ip (cid:54) 90◦ and (Ii,Ωi) are corrected for re-
maining companions, in accord with the invariance relation.
Orbital elements (ai,ei,ωi,Mi) are not fully suitable for tran-
siting systems with small relative inclinations and small eccentric-
ities. To avoid singularities and weakly constrained elements, like
ωi when ei = 0 (circular or weakly eccentric orbits), we use the
Poincar´e modified elements (Xi ≡ ei cosωi, Yi ≡ ei sinωi) instead
of (ei,ωi).
Similarly, the orbital period Pi is more suitable for model fit-
ting than ai since the semi -- major axis depends on the planetary
mass mi (a free parameter) and on the stellar mass m0, which is
fixed to 0.95m(cid:12), but it can be also fitted. Hence, we define Pi as
3 Informations on this project may found at http://git.astri.umk.pl/
projects/mechanic.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
5
one of osculating elements related to ai through the IIIrd Keplerian
law.
The mean anomaly Mi strongly depends on ωi. It determines
the relative orbital phase (the mean longitude) but is also related to
ei. This may be avoided by choosing the time of the first transit Ti
as a free parameter instead of Mi, because it is one of the directly
determined observables from the light-curves. Simple relations be-
tween Ti and Poincar´e canonical elements may be derived easily.
2.6 Direct and indirect transit model parameters
The direct parameters of the transit model are determined from
the basic observables: it is the mean period of transits P∗
i , the
depths, duration times of the transits, and shapes of the light-curves
(through the limb -- darkening coefficients). These data are usually
derived from the period -- phased light -- curves of particular planets.
The depths and durations of transit determine the ratio of planetary
and stellar radii, Ri/R0. If the stellar mass m0 is fixed then Ri and
R0 may be resolved. We can also determine Ii up to the angular mo-
mentum direction invariance, and Ti. In general, the mean period of
transit events P∗
is different from the osculating orbital period at
i
the epoch of the first observation, Pi. A shape of the event -- period
phased light -- curves make it possible to fit the limb darkening co-
efficients, γ1,γ2.
These parameters of the transit model are independent on the
the N-planet dynamics. Hence the remaining are indirect parame-
ters. To resolve them, a dynamical model of the orbital evolution is
required. The indirect parameters consist of planetary masses mi as
well as orbital elements, ei, Ωi, ωi and Pi (instead of P∗
i ). Knowing
mi and Pi, we may fix the osculating semi-major axis ai at the date
of the first (or prescribed) observation.
We would like to note, that the above distinction for two
types of model parameters is somehow arbitrary in our photomet-
ric model. In our algorithm both the direct and indirect parameters
are fitted simultaneously, unlike, for instance, the TTV algorithm,
in which the direct parameters are fitted at the first stage, and the
indirect parameters are fitted in the next step.
Usually, the direct parameters can be estimated much more
reliably than the indirect parameters. Even a potential derivation of
the indirect parameters depend on the particular model of motion.
i.e., kinematic -- Keplerian, or dynamic -- Newtonian, and on the
used method of modelling the observations. In the Keplerian (kine-
matic) model (see, e.g., Agol et al. 2005), mid-transits of a given
planet are governed by geometric reflex motion of the star around
the center of mass in a sub-system composed of the star and all
inner planets. For instance, transits of planet d are affected by plan-
ets b and c, but any outer planet does not affect transits of its inner
companions. Hence, in accord with the Keplerian model, the indi-
rect parameters (mg,ag,eg,Ωg,ωg) of the outermost planet g in the
Kepler-11 system cannot be determined at all.
In a given pair of planets, the outer companion affects the tran-
sits times of the inner planet only through gravitational mutual per-
turbations which lead to changes of osculating orbital elements.
To account for the mutual interactions, one has to apply the self-
consistent N-body model of motion of the system.
Usually, to resolve the indirect parameters from photometric
observations, the well known TTV method is used (Agol et al.
2005). It has two steps. At first, we determine the mean periods, the
mid-transits, and then the (O-C) residua, i.e., differences between
the measured and ephemeris transit times. The (O-C) variations are
observables in the second step during which we search for masses,
eccentricities, and arguments of pericenters of planetary compan-
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
ions. The TTV method in this form has a limitation, because it does
not make any use of transit depths nor their duration times. If the in-
dividual inclinations of planets are different, the planets transit the
parent star usually at different attitudes. Hence the transit depths
as well as duration times may vary, like the (O-C) of mid-transits.
This information can be used to better constrain the transit model.
The mutual inclinations depend on the longitudes of ascending
nodes in accord with
cos∆Ii, j = cosIi cosIj − sinIi sinIj cos(Ωi − Ω j).
Because the inclinations of transiting planets, (Ii, Ij) must be close
to 90◦ then ∆Ii, j ≈ Ωi − Ω j. Within this approximation, the TTV
method is apparently not sensitive for individual Ωi. In fact, differ-
ent values of Ωi imply different mutual inclinations affecting the
dynamics and (O-C). However, the dynamical variability of (O-C)
due to mutual interactions is weaker than the geometric variability
due to changes of transit depths and duration times reflecting the
motion of the star around the mass center of the system.
Overall, by direct modeling of the light-curves (photometric
measurements), rather than the mid-transit times, we can resolve
the (O-C) with an improved precision. Modeling the light-curves
in terms of the N-body model is CPU-demanding, but it makes it
possible to estimate individual longitudes of nodes and mutual in-
clinations. In particular, as will be shown later, the direct method
helped us to derive accurate relative inclinations between planets b
and c, as well as between d and e ∼ 2◦ ± 2◦.
3 RESULTS OF THE BOOTSTRAP ANALYSIS
We performed the direct bootstrap TTV analysis of a few differ-
ent orbital models of the Kepler-11 system. In the most general
case (I), all parameters discussed in the previous section are the free
parameters of the fit model. Some of them are poorly constrained
by the observations, in particular, the eccentricity of planet g and
particular longitudes of nodes. Therefore, we also studied less gen-
eral models, in which some of weakly constrained parameters are
fixed. In the second model (II), Xg = 0, Yg = 0, i.e., eg = 0. In
the third model (III), also Ωg = 0◦, while in the last model (IV),
Ωb,Ωc,Ωd,Ωe,Ωf are all fixed at 0◦. Because inclinations Ii are not
exactly equal to 90◦, also ∆Ii, j (cid:62) 0◦.
For each of these four transit models, we applied the boot-
strap algorithm and we gathered sets of ∼ 1500 solutions for each
instance of the transit model.
eg (cid:54)= 0
3.1 Model I: systems with eg (cid:54)= 0
eg (cid:54)= 0
Figure 2 shows an outcome of the bootstrap algorithm in the form
of normalized histograms constructed for Xg,Yg and Ωg, and de-
picted from the left to the right panel, respectively. The red solid
curves illustrate the best fit asymmetric Gauss function to the his-
togram bins. The formal 1σ errors are marked with red bars dis-
played above the histograms. The best fit parameters correspond-
ing to the maximum of the Gaussian distribution are written in the
respective panels, and they may be compared with the nominal so-
lutions given in Table 1. The uncertainties of the eccentricity and
longitude of node of planet g are relatively large.
Because the nominal system is dynamically unstable, we ex-
amined the whole set of ∼ 1500 bootstrap solutions by calculating
their MEGNO indicator (cid:104)Y(cid:105) on the time interval of ∼ 8000 yr. It
corresponds to ∼ 25,000 periods of the most distant companion.
Such a characteristic time scale should be long enough to detect
6
C. Migaszewski, M. Słonina and K. Go´zdziewski
Figure 2. Bootstrap histograms for Xg,Yg,Ωg, transit model I. See the text for more detail.
unstable solutions due to low -- order 2 -- body and 3 -- body mean mo-
tion resonances (Go´zdziewski et al. 2008, and references therein).
Unfortunately, all initial configurations exhibit large values of (cid:104)Y(cid:105),
indicating that the system is strongly chaotic. The main source of
instability are crossing orbits in the system, that lead to disruptive
events , i.e., one or more of the planets were ejected from the sys-
tem or collided with the parent star. None of the tested solutions
passed the direct integration over 10 Myr.
The parameter space of the Kepler-11 system is ∼ 50-
dimensional, and the dimension of the phase space of the N-body
model is 36-dimensional. The (cid:104)Y(cid:105) experiments indicate that this
system can be locally chaotic and its phase space is filled with
mostly unstable solutions. Then only small regions of stable MMRs
may be present. In the light of a large dimension of the phase
space, the gathered statistics of best -- fit configurations is still very
poor. We conclude that due to short data span of only ∼ 500 days,
and unconstrained elements of the most general model, we can-
not find reliably stable solutions assuming the most general transit
model I. Unfortunately, in this high -- dimensional problem an alter-
nate GAMP algorithm that relies on the optimization with imposed
stability constrains (Go´zdziewski et al. 2008) would be CPU-time
expensive.
3.2 Transit model II: systems with eg = 0
eg = 0
eg = 0
In the next model, we narrow the mostly unconstrained parameters
of the transit model. We fix the eccentricity eg = 0, hence Xg and
Yg are both equal to 0. The results of the bootstrap algorithm are il-
lustrated in Figs. 3-9. All panels in these figures are constructed in
the same manner as Fig. 2. We tested, whether the best -- fit param-
eters encompass at least marginally stable solutions with (cid:104)Y(cid:105) ≈ 2
after T = 16000 yr. Figure 3 shows the normalized histograms for
masses of particular planets expressed in the Earth masses. Besides
formal uncertainties obtained through the bootstrap (filled red cir-
cles), the best-fit parameters derived in (Lissauer et al. 2011) are
plotted (blue filled circles). Clearly, these estimates coincide very
well in both cases. There is one exception though, since the mass
of planet g is not resolved in Lissauer et al. (2011). The direct code
helps to resolve also this mass. It is constrained surprisingly well,
in spite of a narrow observational window. This result confirms our
predictions. Because the orbital model is constrained by all mea-
surements, not the TTVs only, the direct algorithm makes use of
dynamical information contained in the transit depths and widths.
For a reference, black and green asterisks in Fig. 3 mark
masses of the Uranus and Neptune, respectively. The masses of
planets b and f appear in a range specific for the super-Earths. They
are significantly smaller than the masses of two most distant plan-
ets in our Solar system but, as we will show in the next section,
their chemical composition has likely much common with the ice
giants in the Solar system. The next Fig. 4 shows histograms con-
structed for planetary radii expressed in the unit of the Earth radius.
These results confirm data in the discovery paper. Similarly to the
previous plots, the radii of Uranus and Neptune are marked with
asterisks. They are also labeled with RU and RN, respectively. The
derived radius of planet g confirms a hypothesis that it may belong
to the Uranus/Neptune -- class. We note that most of the planets has
radii smaller than RU/N, and only planet e has its radius larger. His-
tograms of the mean densities are presented in Fig. 5. The x-axis is
for the density expressed w.r.t. the Earth density. Black and green
asterisks mark the values characteristic for Uranus and Neptune, re-
spectively. The mean densities of Saturn and water are also marked
with the red and blue symbols, respectively. According to this plot,
the less dense planet e has a density of Saturn. The most dense
planet b may be almost as dense as the Earth. The densities of the
other planets span a range characteristic for Saturn and Neptune,
from ρS to ρN.
Figure 6 is for the bootstrap histograms constructed of the
semi-major axes. These parameters are the best determined among
all of the transit models, with uncertainties of the order of 10−5 au
only. We do not compare these results with data in (Lissauer et al.
2011) because they accounted for the formal error of the stellar
mass. Note that we fixed m0 = 0.95m(cid:12), because we found that this
parameter is unconstrained by the photometric data. Yet it seems
that the χ2
ν(m0) function monotonically increases in the range of
m0 ∈ (0.7,1.2)m(cid:12).
The first five panels of Fig. 7 are for the eccentricities, and the
bottom, right-hand panel is for ∆ωb,c ≡ ωb −ωc . These histograms
confirm that the eccentricities of planets b to f are small, typically
less that 0.05, and the arguments of pericenters are not well con-
strained. The last panel assures us that ∆ωb,c is determined with an
error of only ∼ 10◦, recalling a narrow time -- window of the photo-
metric data. The best -- fit parameters of model II are given in Tab. 2.
Inclination Ib was constrained to the (cid:54) 90◦ range, and due
to the invariance rule implied by the direction of the total angular
momentum, the remaining inclination Ii may be smaller and larger
than 90◦. We tested whether there is a correlation of the transit
events with a given half -- disc of the star. We found that both cases
are equally possible. Because the orbits are inclined to the plane
of the sky at angles close to 90◦, the relative inclinations with the
same longitudes of nodes may be ∼ 2◦ -- 3◦. As expected, the in-
direct parameters Ωi are unconstrained, see Tab. 2. Therefore, the
main contribution to the uncertainties of the relative inclinations
comes from ambiguous estimates of Ωi rather than of Ii.
Curiously, there appears a clear correlation between mutual
inclinations in particular pairs of orbits, namely c and e, f and e,
as well as d and f. This can be seen in normalized histograms
constructed for the inclinations, Fig. 8. For a chosen planet, we
transform Ii to (cid:54) 90◦ range (in accord with the inclination in-
variance rule), and we compute the bootstrap histogram for Ij.
Panels of Fig. 8, from the left to the right, are for pairs (i, j) =
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
eg cos ωg = -0.26 +0.16−0.08 0 0.02 0.04 0.06 0.08-0.3-0.2-0.1 0 0.1 0.2 0.3eg sin ωg = 0.043 +0.058−0.085 0 0.02 0.04 0.06 0.08 0.1-0.3-0.2-0.1 0 0.1 0.2 0.3Ωg = 7 +42−53 deg 0 0.02 0.04 0.06 0.08-180-135-90-45 0 45 90 135 180A dynamical analysis of the Kepler-11 planetary system
7
Table 1. Bootstrap results for transit model I. Mass of the star is 0.95m(cid:12) (fixed). The best-fitting stellar parameters of this model are R0 = 1.140+0.030
−0.027,
γ1 = 0.33+0.47−0.30, γ2 = 0.41+0.24−0.34, γ1 + γ2 = 0.74+0.23−0.23. Osculating Poincar´e elements are given at the epoch of the first observation JD 2455964.51128.
parameter/planet
b
c
d
e
f
g
18+24−15
3.80+0.15−0.14
0.33+0.47−0.23
(cid:17)
(cid:16)+59−28
0.463918
−0.26+0.16−0.08
0.008+0.058
−0.085
90.23+0.16−0.11
−65+51−46
6+7−19
118.410
(cid:16)+16−10
501.916(cid:0)+42−21
(cid:17)
(cid:1)
m [m⊕]
R [R⊕]
¯ρ [¯ρ⊕]
a [au]
e cosω
e sinω
I∗ [deg]
Ω [deg]
M + ω [deg]
P [d]
T0 [JD]
4.2+2.4−3.0
2.04+0.18−0.10
0.50+0.38−0.29
0.091089(cid:0)+13−11
(cid:1)
0.010+0.017
−0.021
−0.011+0.031
−0.025
88.40+0.76−0.42
0 (fixed)
204.5+2.2−2.5
10.3023(cid:0)+24−18
471.505(cid:0)+20−7
(cid:1)
(cid:1)
9.2+3.8−6.9
3.25+0.13−0.14
0.27+0.14−0.17
0.106522(cid:0)+7−12
(cid:1)
0.005+0.017
−0.018
−0.004+0.028
−0.020
91.17+0.40−0.20
3.2+4.2−2.9
265.3+2.0−2.0
13.0284(cid:0)+12−20
471.175(cid:0)+20−4
(cid:1)
(cid:1)
(cid:1)
8.9+3.5−2.7
3.58+0.17−0.14
0.19+0.07−0.07
0.154241(cid:0)+19−10
−0.013+0.008
−0.022
−0.009+0.006
−0.015
89.18+0.22−0.17
−33+13−11
182.8+2.3−1.0
22.7002(cid:0)+41−24
(cid:16)+16−6
481.455
(cid:1)
(cid:17)
10.7+2.4−2.1
4.71+0.20−0.18
0.10+0.04−0.02
(cid:17)
(cid:16)+15−21
0.193937
−0.020+0.008
−0.022
−0.016+0.007
−0.011
88.743+0.062
−0.060
−31+12−11
197.4+1.8−1.4
32.0051(cid:0)+42−46
487.178(cid:0)+22−8
(cid:1)
(cid:1)
(cid:1)
3.6+5.4−2.6
2.82+0.19−0.14
0.16+0.22−0.14
0.249489(cid:0)+39−25
−0.006+0.011
−0.018
−0.017+0.016
−0.021
89.30+0.12−0.09
−32+30−27
89.5+1.5−1.8
(cid:1)
46.700(cid:0)+11−6
(cid:16)+15−8
(cid:17)
464.673
Figure 3. Bootstrap histograms for planetary masses, transit model II.
Figure 4. Bootstrap histograms for the planetary radii, transit model II.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
mb = 4.2 +2.8−2.4 mE∗U 0 0.02 0.04 0.06 0.08 0 3 6 9 12 15mc = 9.0 +3.6−6.0 mE∗∗NU 0 0.02 0.04 0.06 0.08 0 5 10 15 20 25md = 8.0 +3.5−2.6 mE∗∗NU 0 0.03 0.06 0.09 0.12 0 5 10 15 20 25me = 10.2 +2.4−1.0 mE∗∗NU 0 0.04 0.08 0.12 0.16 0 5 10 15 20 25mf = 3.5 +4.7−1.8 mE∗∗NU 0 0.03 0.06 0.09 0.12 0 6 12 18 24 30mg = 16 +16−11 mE∗∗NU 0 0.03 0.06 0.09 0.12 0 20 40 60 80 100 120Rb = 2.02 +0.16−0.13 RE 0 0.02 0.04 0.06 0.08 1.6 1.8 2 2.2 2.4 2.6Rc = 3.18 +0.16−0.12 RE 0 0.02 0.04 0.06 0.08 2.9 3.1 3.3 3.5 3.7Rd = 3.59 +0.17−0.14 RE∗∗NU 0 0.02 0.04 0.06 0.08 0.1 3.2 3.4 3.6 3.8 4 4.2Re = 4.53 +0.30−0.19 RE∗∗NU 0 0.02 0.04 0.06 0.08 0.1 4 4.2 4.4 4.6 4.8 5 5.2 5.4Rf = 2.78 +0.17−0.15 RE 0 0.02 0.04 0.06 0.08 0.1 2.3 2.5 2.7 2.9 3.1 3.3Rg = 3.78 +0.20−0.13 RE∗∗NU 0 0.04 0.08 0.12 0.16 2.4 2.6 2.8 3 3.2 3.4 3.6 3.8 4 4.2 4.4 4.68
C. Migaszewski, M. Słonina and K. Go´zdziewski
Figure 5. Bootstrap histograms for the mean densities, transit model II.
Figure 6. Bootstrap histograms for the semi-major axes, transit model II.
Figure 7. Bootstrap histograms for eccentricities and ∆ωb,c, transit model II.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
〈ρb〉 = 0.42 +0.42−0.30 〈ρE〉∗∗∗∗SNU 0 0.02 0.04 0.06 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8〈ρc〉 = 0.27 +0.14−0.23 〈ρE〉∗∗∗∗SNUH2O 0 0.02 0.04 0.06 0.08 0 0.2 0.4 0.6 0.8〈ρd〉 = 0.17 +0.07−0.07 〈ρE〉∗∗∗∗SNUH2O 0 0.02 0.04 0.06 0.08 0.1 0 0.08 0.16 0.24 0.32 0.4 0.48 0.56〈ρe〉 = 0.11 +0.03−0.03 〈ρE〉∗∗∗∗SNUH2O 0 0.03 0.06 0.09 0.12 0 0.08 0.16 0.24 0.32〈ρf〉 = 0.17 +0.21−0.14 〈ρE〉∗∗∗∗SNU 0 0.03 0.06 0.09 0.12 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6〈ρg〉 = 0.26 +0.30−0.16 〈ρE〉∗∗∗∗ 0 0.03 0.06 0.09 0.12 0 0.4 0.8 1.2 1.6 2 2.4 2.8 3.2ab = 0.091088(+15 −7) au 0 0.02 0.04 0.06 0.08 0.1 0.09104 0.09106 0.09108 0.0911 0.09112 0.09114ac = 0.106523(+6−11) au 0 0.02 0.04 0.06 0.08 0.1 0.10649 0.10651 0.10653 0.10655ad = 0.154242(+16−10) au 0 0.02 0.04 0.06 0.08 0.1 0.1542 0.15423 0.15426 0.15429 0.15432ae = 0.193941(+16−17) au 0 0.03 0.06 0.09 0.12 0.19386 0.1939 0.19394 0.19398 0.19402af = 0.249514(+32−23) au 0 0.04 0.08 0.12 0.16 0.24934 0.24942 0.2495 0.24958 0.24966 0.24974ag = 0.46398(5) au 0 0.05 0.1 0.15 0.2 0.4635 0.4637 0.4639 0.4641 0.4643eb = 0.002 +0.023−0.001 0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08ec = 0.003 +0.017−0.002 0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06ed = 0.002 +0.010−0.002 0 0.02 0.04 0.06 0.08 0.1 0 0.01 0.02 0.03 0.04ee = 0.002 +0.010−0.002 0 0.02 0.04 0.06 0.08 0.1 0 0.01 0.02 0.03 0.04ef = 0.006 +0.015−0.004 0 0.04 0.08 0.12 0.16 0 0.03 0.06 0.09 0.12∆ωb,c = -1 +8−13 deg 0 0.05 0.1 0.15 0.2-120-75-30 15 60 105A dynamical analysis of the Kepler-11 planetary system
9
Table 2. Bootstrap results for model II (with fixed eg = 0). Mass of the star is 0.95m(cid:12) (fixed). Best fitted stellar parameters are R0 = 1.161+0.035
−0.028,
γ1 = 0.32+0.46−0.30, γ2 = 0.41+0.17−0.36, γ1 + γ2 = 0.73+0.16−0.30. Osculating Poincar´e elements are given at the epoch of the first observation JD 2455964.51128.
parameter/planet
b
c
d
e
f
m [m⊕]
R [R⊕]
¯ρ [¯ρ⊕]
a [au]
e cosω
e sinω
I∗ [deg]
Ω [deg]
M + ω [deg]
P [d]
T0 [JD]
4.2+2.8−2.4
2.07+0.16−0.13
0.48+0.42−0.30
(cid:17)
(cid:16)+15−7
0.091087
0.006+0.011
−0.022
0.020+0.027
−0.027
88.39+0.95−0.24
0 (fixed)
205.0+2.2−2.3
10.3019(cid:0)+23−14
471.504(cid:0)+21−7
(cid:1)
(cid:1)
9.2+3.6−6.0
3.31+0.16−0.12
0.25+0.14−0.23
(cid:17)
(cid:16)+6−11
0.106521
0.001+0.010
−0.017
0.024+0.025
−0.023
91.17+0.38−0.22
3.4+4.3−2.7
265.7+2.1−1.2
13.0281(cid:0)+14−18
471.176(cid:0)+18−4
(cid:1)
(cid:1)
9.2+3.5−2.6
3.65+0.17−0.14
0.19+0.07−0.07
(cid:17)
(cid:16)+16−10
0.154233
−0.013+0.007
−0.019
0.008+0.008
−0.015
89.14+0.23−0.13
−23+12−12
182.7+1.9−0.9
22.6985(cid:0)+34−22
481.454(cid:0)+14−6
(cid:1)
(cid:1)
10.5+2.4−1.0
4.80+0.30−0.19
0.10+0.03−0.03
(cid:17)
(cid:16)+16−17
0.193926
−0.020+0.007
−0.015
−0.002+0.005
−0.014
88.701+0.052
−0.076
−22+11−12
197.4+1.6−1.0
(cid:16)+45−34
487.177(cid:0)+19−9
32.0025
(cid:17)
(cid:1)
4.4+4.7−1.8
2.88+0.17−0.15
0.19+0.21−0.14
0.249511(cid:0)+32−23
−0.004+0.016
−0.013
−0.007+0.020
−0.016
89.282+0.09−0.11
−25+24−27
89.2+1.7−1.5
46.7054(cid:0)+89−57
464.670(cid:0)+14−9
(cid:1)
(cid:1)
g
3+16−3
3.93+0.20−0.13
0.04+0.30−0.04
0.463924(cid:0)+42−37
(cid:1)
(cid:1)
0 (fixed)
0 (fixed)
90.31+0.086
−0.055
37+63−59
336.282+0.049
−0.093
(cid:1)
118.4147(cid:0)+90−33
(cid:1)
501.916(cid:0)+40−11
Figure 8. Bootstrap histograms for absolute inclinations, transit model II.
(e,c), (e,f), (f,d). If Ie (cid:54) 90◦ then much more likely Ic,If (cid:54) 90◦
than Ic,If (cid:62) 90◦. Similarly, if If (cid:54) 90◦, then Id (cid:54) 90◦ appears more
likely than If (cid:62) 90◦.
For particular pairs of planets, the relative inclinations can
be determined surprisingly well. Figure 9 shows the bootstrap his-
tograms ∆Ii, j for such pairs which exhibit well constrained values.
These histograms reveal that orbits of planets b and c are almost
coplanar. Similarly, the pair of planets d and e form an almost
coplanar sub-system. The mutual inclinations of orbits in these
pairs are less than 5◦, with most likely values of 2◦ -- 3◦. The re-
maining panels indicate that the mutual inclinations between five
inner orbits remain within a few degrees range. Their upper limits
are not so small as in the first two sub-systems. The outermost orbit
of planet g may by highly inclined to the rest of the system, see
errors of Ωg in Tab. 2.
These results confirm a hypothesis in the discovery paper. In
accord with this work, planetary orbits in the Kepler -- 11 system
should be mutually inclined by no more than a few degrees. It flows
from estimating a probability that for a given orientation of the or-
bits, all six planets transit the star. This reasoning assumes that all
inclinations are independent. However, we found that Kepler -- 11
system is composed of two or three sub-systems, which exhibit
small mutual inclinations of orbits. Although a probability that
the mutual inclinations between these sub-systems are significant
seems a bit larger than for fully independent orbits, it still remains
very small. We estimate that a randomly located observer can de-
tect transits of all 6 planets with a probability as small as ∼ 0.05%,
for both models I and II.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
eg = 0,Ωg = 0
eg = 0,Ωi = 0
eg = 0,Ωg = 0) and IV (eg = 0,Ωi = 0
3.3 Models III (eg = 0,Ωg = 0
eg = 0,Ωi = 0)
The results for model III and model IV are given in Tabs. 3 and
4. Most of these results are common for all transit models I to IV.
There are some differences regarding a determination of the mass
of planet g. In the realm of models III and IV (note that both have
fixed eg = 0 and Ωg = 0), only an upper limit of mg smaller than
20−30 Earth masses may be found. The low limits of mg in model I
are likely due to weakly constrained eg and Ωg.
Let us recall that in the bootstrap set derived for model II, we
found only two solutions with (cid:104)Y(cid:105) ≈ 2 after 16000 yr. However,
this integration time scale is too short to detect weak instability
which actually leads to catastrophic disruption of these configura-
tions. It was verified by the direct, long -- term integrations. Hence,
we did not detect any long -- term stable configuration in the boot-
strap set of model II. Similarly, stability tests performed for config-
urations of model III did not reveal any stable models. As compared
to model II, fixed Ωg = 0 seem does not change the general view of
the stability of the system.
For model IV we found many stable configurations which con-
firm stability analysis in Lissauer et al. (2011). They found some
stable solutions assuming that the Kepler-11 system is strictly co-
planar. We may conclude that a factor of small relative inclinations
is more important for maintaining the long term stability than small
eccentricities. This will be discussed in more detail further in this
work.
We examined a probability that a randomly located observer
could detect transits of all planets in the system. This is basically
unlikely for model III (∼ 0.09%), while for model IV a probability
of such an event is larger, and we estimate it ∼ 3.4%.
Ic = 88.70 +0.40−0.22 degIc = 91.07 +0.25−0.31 degIe ≤ 90 deg 0 0.03 0.06 0.09 88 89 90 91 92If = 89.34 +0.09−0.11 degIf = 90.74 +0.07−0.10 degIe ≤ 90 deg 0 0.04 0.08 0.12 0.16 89 89.5 90 90.5 91Id = 89.09 +0.21−0.12 degId = 90.76 +0.18−0.15 degIf ≤ 90 deg 0 0.04 0.08 0.12 88.5 89 89.5 90 90.5 91 91.510
C. Migaszewski, M. Słonina and K. Go´zdziewski
Table 3. Bootstrap results for model III (eg = 0,Ωg = 0). Mass of the star is 0.95m(cid:12) (fixed). Fitted stellar parameters: R0 = 1.158+0.021
γ2 = 0.41+0.25−0.40, γ1 + γ2 = 0.73+0.32−0.21. Osculating Poincar´e elements are given at the epoch of the first observation JD 2455964.51128.
−0.038, γ1 = 0.32+0.44−0.25,
parameter/planet
b
c
d
e
f
m [m⊕]
R [R⊕]
¯ρ [¯ρ⊕]
a [au]
e cosω
e sinω
I∗ [deg]
Ω [deg]
M + ω [deg]
P [d]
T0 [JD]
4.0+2.5−3.0
2.07+0.13−0.13
0.45+0.40−0.33
(cid:17)
(cid:16)+15−9
0.091088
0.015+0.013
−0.025
0.007+0.034
−0.019
88.35+0.83−0.45
0 (fixed)
204.0+2.8−1.7
(cid:16)+25−15
471.504(cid:0)+20−9
10.3020
(cid:17)
(cid:1)
9.1+3.3−6.1
3.30+0.13−0.13
0.25+0.12−0.17
0.106519(cid:0)+9−10
(cid:1)
0.008+0.015
−0.017
0.012+0.025
−0.022
88.80+0.38−0.20
−3.5+4.2−2.5
264.9+1.9−1.8
(cid:16)+16−18
471.176(cid:0)+17−4
13.0278
(cid:17)
(cid:1)
(cid:1)
9.1+3.5−2.9
3.64+0.14−0.16
0.19+0.09−0.07
0.154234(cid:0)+22−6
−0.008+0.010
−0.016
0.000+0.007
−0.013
90.88+0.20−0.14
21+14−12
182.2+1.5−1.2
22.6986(cid:0)+43−18
481.454(cid:0)+17−6
(cid:1)
(cid:1)
(cid:1)
10.6+2.9−2.1
4.79+0.19−0.20
0.10+0.05−0.02
0.193924(cid:0)+24−13
−0.015+0.009
−0.011
−0.010+0.005
−0.013
88.701+0.071
−0.066
20+14−11
196.7+1.3−1.2
32.0019(cid:0)+49−34
487.176(cid:0)+21−6
(cid:1)
(cid:1)
(cid:1)
4.3+5.8−2.7
2.88+0.17−0.15
0.18+0.27−0.09
0.249511(cid:0)+31−23
−0.002+0.012
−0.015
−0.011+0.016
−0.023
89.26+0.10−0.12
22+24−23
89.0+1.9−1.6
46.7056(cid:0)+87−55
464.671(cid:0)+13−10
(cid:1)
(cid:1)
g
1+28−1
3.92+0.12−0.15
0.02+0.49−0.04
(cid:17)
(cid:16)+55−36
0.463917
0 (fixed)
0 (fixed)
89.699+0.077
−0.064
0 (fixed)
336.29+0.05−0.13
118.412
(cid:16)+26−12
501.915(cid:0)+43−14
(cid:17)
(cid:1)
Figure 9. Bootstrap histograms for the mutual inclinations, transit model II.
4 DISCUSSION ON THE PLANET INTERIORS
Figure 10 shows bootstrap diagrams of a few selected pairs of pa-
rameters. These results are for model II. The top row is for the semi-
major axes and the planetary masses, the radii and mean densities,
respectively. The red and green curves mark the data for Uranus and
Neptune. The bottom row is for the mass -- radius, mass -- density and
radius -- density relations, respectively. Similarly, the red and green
filled circles are for Uranus and Neptune. As we concluded above,
the orbital solutions in set II are only marginally stable, however, it
is a matter of unconstrained orbital angles. Note that a discussion
in this Section concerns semi-major axes (known with an excellent
precision) as well as planetary masses and radii.
This figure reveals that the most inner four planets in the
Kepler-11 system exhibit a clear and curious anti/correlation of
masses, radii and densities with the semi-major axes. Masses and
radii increase with ai, while densities decrease. The last panel con-
structed for (R,ρ) shows a weak anti-correlation between the radii
and densities, the smaller radius, the larger density. The determined
masses and radii of the planets provide some insight into their
chemical composition. We use a simple analytic relation between
the radius and the mass of a cold body (Lynden-Bell & O'Dwyer
2001; Lynden-Bell & Tout 2001) to estimate the characteristic den-
sity ρ0 of planetary matter. This density can be compared with ρ0
calculated for a given number of nucleons per number of electrons
of a chemical mixture forming the planet, µe. The value of µe is a
simple mean over the elements in each chemical substance or com-
ponent. For instance the H-He mixture has µe = 8/7 for the mass
proportions 3 to 1, and ice or rock has µe ≈ 2. In this way, we can
obtain some insight into likely chemical composition of the planets.
Our results for Kepler-11 are presented in Fig. 11. Black
curves with grey areas are for ρ0 and its uncertainty ∆ρ0. Each
panel is for one planet of the Kepler-11 system. Data for planets b
to planet g are displayed from the left to the right, respectively.
The colored curves are for the Solar system, i.e., for Uranus (red),
Neptune (green), Jupiter (blue), Saturn (violet) and the Earth (light
blue). The density ρ0 was computed in a wide range of µe ∈ [1,2].
These values are known relatively well for the Sun companions.
Following Helled et al. (2011), for Uranus and Neptune one finds
µe ≈ 1.1 (for the icy model) and µe ≈ 1.35 (for the rocky model).
The density ρ0 in these particular case is plotted with filled cir-
cles. Similarly, for Jupiter and Saturn, µe may be also estimated
≈ 1.08− 1.09 (Guillot 1999). Values of ρ0 for these particular µe
are marked with circles. Let us note that densities ρ0 of Jupiter and
Saturn are almost identical.
Lynden-Bell & Tout (2001) argue that ρ0 is the zero -- pressure
density ρ0,p of the chemical mixture of the planets. Because their
model has many simplifications, ρ0 is usually 2 − 5 times larger
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
∆Ib,c = 2.7 +2.4−1.1 deg 0 0.03 0.06 0.09 0.12 0 5 10 15 20∆Id,e = 2.3 +2.4−1.1 deg 0 0.03 0.06 0.09 0.12 0 5 10 15 20∆Ie,f = 1.9 +16.2−1.4 deg 0 0.02 0.04 0.06 0.08 0.1 0 30 60 90∆Ib,e = 2.4 +8.6−1.1 deg 0 0.02 0.04 0.06 0.08 0.1 0 15 30 45∆Ic,e = 0.8 +14.0−0.7 deg 0 0.02 0.04 0.06 0.08 0 10 20 30 40 50∆Ic,d = 1.4 +13.9−0.9 deg 0 0.02 0.04 0.06 0.08 0 10 20 30 40 50A dynamical analysis of the Kepler-11 planetary system
11
Table 4. Bootstrap results for model IV (eg = 0,Ωi = 0,i = b,c,d,e, f ,g). Mass of the star is 0.95m(cid:12) (fixed). The best -- fit stellar parameters: R0 = 1.140+0.026
−0.024,
γ1 = 0.30+0.32−0.30, γ2 = 0.42+0.18−0.42, γ1 + γ2 = 0.72+0.22−0.25. Osculating Poincar´e elements are given at the epoch of the first observation JD 2455964.51128.
parameter/planet
b
c
d
e
f
g
m [m⊕]
R [R⊕]
¯ρ [¯ρ⊕]
a [au]
e cosω
e sinω
I∗ [deg]
M + ω [deg]
P [d]
T0 [JD]
(cid:1)
3.3+2.4−1.8
2.01+0.13−0.13
0.40+0.33−0.19
0.091088(cid:0)+13−11
−0.002+0.015
−0.021
0.049+0.020
−0.050
88.76+0.96−0.41
205.8+2.4−2.1
10.3021(cid:0)+24−18
471.505(cid:0)+20−7
(cid:1)
(cid:1)
8.8+4.0−5.0
3.23+0.12−0.12
0.26+0.13−0.18
(cid:17)
(cid:16)+6−8
0.106514
−0.006+0.015
−0.016
0.050+0.020
−0.044
91.00+0.40−0.25
266.4+1.7−1.9
13.0269(cid:0)+10−17
471.177(cid:0)+18−3
(cid:1)
(cid:1)
(cid:1)
8.9+2.3−3.4
3.59+0.15−0.13
0.19+0.07−0.08
0.154234(cid:0)+18−7
−0.012+0.009
−0.019
−0.010+0.008
−0.018
90.89+0.17−0.18
182.7+2.1−1.2
22.6986(cid:0)+43−13
481.452(cid:0)+17−6
(cid:1)
(cid:1)
(cid:1)
10.3+1.9−1.5
4.70+0.21−0.15
0.10+0.03−0.03
0.193927(cid:0)+19−10
−0.018+0.006
−0.017
−0.016+0.006
−0.017
88.743+0.043
−0.049
197.2+1.8−0.9
32.0027(cid:0)+41−30
487.176(cid:0)+22−6
(cid:1)
(cid:1)
4.1+3.8−2.4
2.82+0.15−0.15
0.18+0.20−0.10
(cid:17)
(cid:16)+35−26
0.249507
−0.007+0.015
−0.020
−0.017+0.011
−0.021
89.30+0.10−0.09
89.6+2.6−1.3
46.7044(cid:0)+88−59
464.670(cid:0)+14−9
(cid:1)
(cid:1)
< 21
3.85+0.12−0.12
< 0.35
(cid:17)
(cid:16)+57−39
0.463913
0 (fixed)
0 (fixed)
89.719+0.061
−0.068
336.286+0.066
−0.057
118.411(cid:0)+14−15
(cid:16)+26−11
501.914
(cid:1)
(cid:17)
Figure 10. The top row: mass, radius and mean density as a function of the semi-major axis (model II). Black and green solid curves are for Uranus and
Neptune, respectively. The bottom row: mass -- radius, mass -- mean density and radius -- mean density relations.
then ρ0,p. Keeping this in mind, the densities ρ0 calculated for
Kepler-11 planets can be compared with those of the Solar sys-
tem planets. The value of ρ0 is best determined for planet e. Its
is very close to the Jupiter/Saturn (J/S) value ∼ 0.45gcm−3. This
suggests, that planet e is built mainly of a H/He mixture with mass
proportions of the elements close to 3/1 with a portion of heavier
elements contained in ices or rocks. This makes the planet classified
as a smaller "cousin" of Jupiter and Saturn rather than of Neptune
and Uranus, as suggested in (Lissauer et al. 2011).
The density ρ0 of planet b is determined worse than for
planet e. It is rather unlikely though that it belongs to the same
class as planet e. Parameter ρ0 is larger than for Jupiter and Saturn,
even taking into account a large uncertainty. It is also larger than ρ0
for Uranus and Neptune (U/N) -- like planets but is smaller than ρ0
for the Earth. We can conclude that planet b is a small planet con-
taining a large percentage of heavy elements in its interior, which
is likely larger than in the ice giants. It is reasonable to classify this
planet in the super-Earths family, although, it may be also a small
Neptune -- like planet.
Planet f has the mass similar to planet b. However,
its
composition is likely between the Jupiter -- Saturn and Uranus --
Neptune classes. Planet d has likely similar composition as planet f.
The best-fit estimate of ρ0 for planet c is very close to the
Uranus/Neptune value. For planet g there is only the upper limit
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
of ρ0. However, it is probably close to the Jupiter/Saturn value or,
less likely, to the Neptune/Uranus value.
These conclusions are reinforced after inspecting the bottom-
left panel of Fig. 10 illustrating the mass -- radius diagram for the
Kepler-11 system. The mass -- radius relation computed for ρ0 and
µe of the Solar system planets are plotted with different colors. Data
are shown for Uranus (red), Neptune (green), Jupiter (blue), Saturn
(violet) and Earth (light blue), respectively. Representations of this
relation are plotted in the mass-density diagram (the middle panel)
and the radius -- density diagram (the right panel).
For small masses, the characteristic density ρ0 is very close to
ρ (see Eq. 34 in Lynden-Bell & O'Dwyer 2001). This derived de-
cay of ρ with the mean distance from the star suggests that the inner
planets may contain larger amount of heavy elements than more
distant companions. If this correlation can be confirmed, it may
provide an observational constraint for the planet formation theory.
Allowing for some speculations here, let us note that all Kepler-11
planets exhibit masses in the same range of a few Earth masses.
Hence, they likely have formed in a similar way and physical envi-
ronment (Rogers et al. 2011). Small eccentricities and small relative
inclinations suggest that the system evolved orbitally smoothly to-
wards the current state, conserving the ordering of initial distances
from the star. The observed relation between ρ0 and ai may then
a [au] : m [mE] 0 5 10 15 20 0.08 0.18 0.28 0.38 0.48a [au] : R [RE] 1.8 2.6 3.4 4.2 5 0.08 0.18 0.28 0.38 0.48a [au] : ρ [ρE] 0 0.2 0.4 0.6 0.8 1 0.08 0.18 0.28 0.38 0.48 1.8 2.6 3.4 4.2 5 0 5 10 15 20m [mE] : R [RE]bcdefgUN 0 0.2 0.4 0.6 0.8 1 0 5 10 15 20m [mE] : ρ [ρE]bcdefgUN 0 0.2 0.4 0.6 0.8 1 1.8 2.6 3.4 4.2 5R [RE] : ρ [ρE]bcdefgUN12
C. Migaszewski, M. Słonina and K. Go´zdziewski
Figure 11. Characteristic density ρ0 of the chemical mixture of planetary interiors as functions of the mean number of nucleons per one electron, µe (model II).
indicate the chemical composition and mass density distribution in
the primordial protoplanetary disk.
We underline that the results in this section must be consid-
ered as preliminary. Due to relatively narrow time -- window of the
photometric data, masses of the planets are determined with large
uncertainties.
5 RESULTS OF THE DYNAMICAL ANALYSIS
The best -- fit solutions gathered with the help of the bootstrap al-
gorithm provide us primary information required to perform exten-
sive study of the dynamical stability of the system. Because the or-
bits of Kepler-11 super-Earth planets are confined within the mean
distance of Mercury in the Solar system, we could expect that the
dynamics of this system are extremely complex. Indeed, prelim-
inary integrations demonstrated that the Kepler-11 system is dy-
namically packed, in accord with a definition and the PPS hypoth-
esis in (Barnes et al. 2008). In spite of apparently ordered config-
urations with quasi -- circular, almost coplanar orbits, and relatively
small masses, no long-term stable model I solutions were found.
Below, we try to resolve this paradox and we try to detect sources of
this seemingly odd and strong instability. To illustrate the structure
of the phase space close to the best -- fit configurations, we choose
a few representative solutions and we construct the MEGNO maps
in their vicinity.
5.1 Quasi-stable solutions in transit model II
For model II, among ∼ 1500 initial conditions, we found only 2
configurations exhibiting MEGNO close to 2 after T = 16000 yr.
Parameters of these solutions are listed in Tabs. 5 and 6. The first
stable solution (refereed to as IIa from hereafter, see Tab. 5), has a
relatively low mass of planet c ∼ 1.5 Earth masses. Two innermost
planets b and c have almost coplanar orbits. The next three planets,
d, e and f also form a nearly coplanar sub-system (d-e-f), which
is inclined to the first two orbits by large angle ∼ 15◦. The outer-
most orbit is inclined even more, by ∼ 50◦ to the inner subsystem
of (b-c), and by ∼ 30◦to the triple -- planet subsystem of (d-e-f). The
second stable solution IIb (see Tab. 6) has all masses close to the
nominal best -- fit values. The mutual inclinations of five inner or-
bits are close to 0◦, while the outermost orbit of planet g is highly
inclined to the inner orbits by ∼ 90◦, similarly to solution IIa. Be-
cause the relative inclinations between particular pairs of orbits are
large in these best-fit solutions, such systems might be unlikely ob-
served by a randomly located observer. We estimate a probability
of such an event as ∼ 0.07% and ∼ 0.05% for solutions IIa and IIb,
respectively.
5.2 Triple-planet resonances as the main source of instability
Let us now study the vicinity of these particular solutions through
the dynamical maps. For each initial condition of the discrete grid
with 512× 512 resolution, we compute (cid:104)Y(cid:105) over T = 8000 yr. Fig-
ure 12 shows the MEGNO maps for solution IIa. Each panel is for
a different pair of planets. The coordinate axes are rescaled mean
motions centered at their nominal ni,0:
xi ≡ ni − ni,0
ni,0
× 104.
The xi -- axes span 1σ uncertainties of the semi -- major axes ai, in
accord with Tab. 2. The semi-major axes are determined very pre-
cisely, hence the 1σ interval span a range of 10−5 to 10−4 au. The
rescaled xi are confined to order of 10.
The MEGNO is color-coded in the dynamical maps. Blue re-
gions mean regular solutions with (cid:104)Y(cid:105) ≈ 2, while yellow color is
for chaotic (unstable) initial conditions, (cid:104)Y(cid:105) (cid:62) 5. The resolution is
512× 512 pixels, total integration time is T = 8000 yr per pixel,
SABA4 integrator step-size is 0.5 days. A single computation of
each map took ∼ 16 hrs of 1200 CPU cores. The integrations of
each pixel were performed up to the end time T , regardless a value
of MEGNO.
Still, although the maps cover tiny regions of the phase space,
close to the fixed initial condition, they reveal a sophisticated struc-
ture. Because we consider the dynamics in terms of conservative,
close to integrable Hamiltonian system, this structure is governed
by the resonant motions. A relatively short integration time ∼ 104 --
105 characteristic periods, equivalent to the orbital period of the
outermost planet makes it possible to detect unstable MMRs. They
appear as yellow straight bands of different widths and slopes.
Inspecting the dynamical maps, we can identify particular reso-
nances.
A condition for the mean motion resonance in the N-planet
system may be written in the following form:
N
∑
i=1
pi
dλi
d t = O [ fω, fΩ] ,
or
N
∑
i=1
pini = O [ fω, fΩ] ,
(3)
where ni is the mean motion of the i-th planet, fω and fΩ are the
fundamental frequencies associated with the pericenter arguments
ωi and the longitudes of nodes Ωi (for all orbits), and pi are rela-
tively prime integers. The linear relations must obey the d'Alambert
rule.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
0123456µeρ0 [g/cm3]bcdefg 1 1.2 1.4 1.6 1.8µebcdefg 1 1.2 1.4 1.6 1.8µebcdefg 1 1.2 1.4 1.6 1.8µebcdefg 1 1.2 1.4 1.6 1.8µebcdefg 1 1.2 1.4 1.6 1.8µebcdefg 1 1.2 1.4 1.6 1.8A dynamical analysis of the Kepler-11 planetary system
13
Figure 12. Symplectic MEGNO maps in the (xi,x j) -- plane for solution IIa (see the text for details). Color scale for (cid:104)Y(cid:105) is [1,5] (blue means (cid:104)Y(cid:105) ≈ 2 and stable
solutions; yellow is for (cid:104)Y(cid:105) (cid:38) 5 and unstable motions). Each panel is for different pair of planets labeled in its bottom-right corner.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
14
C. Migaszewski, M. Słonina and K. Go´zdziewski
Table 5. Orbital parameters of marginally stable configuration IIa. Mass of the star is 0.95m(cid:12). Osculating Poincar´e elements are given at the initial epoch
JD 2455964.51128.
parameter/planet
b
c
d
e
f
g
m [m⊕]
a [au]
e
I [deg]
Ω [deg]
ω [deg]
M [deg]
4.550
0.091097
0.00800
89.048
0 (fixed)
158.534
47.28591
1.542
0.106515
0.00698
88.938
2.119
138.150
128.23248
7.224
0.154241
0.00918
89.023
15.216
61.831
118.94929
15.698
0.193939
0.00924
88.803
14.804
183.853
12.27874
4.340
0.249532
0.01827
89.339
20.626
296.192
151.62481
18.530
0.463815
0 (fixed)
89.470
52.574
0 (fixed)
336.22407
Table 6. Orbital parameters of solution IIb. Mass of the star is 0.95m(cid:12). Osculating Poincar´e elements are given at the initial epoch JD 2455964.51128.
parameter/planet
b
c
d
e
f
g
m [m⊕]
a [au]
e
I [deg]
Ω [deg]
ω [deg]
M [deg]
6.227
0.091102
0.02314
88.000
0 (fixed)
178.847
29.83704
4.422
0.106525
0.01780
90.849
−1.748
175.740
92.04062
8.467
0.154254
0.01148
89.296
−5.944
35.793
144.32568
11.293
0.193942
0.00401
91.206
−2.902
336.443
218.20059
6.866
0.249501
0.01859
90.677
−2.393
350.572
96.06246
32.651
0.464000
0 (fixed)
89.733
92.907
0 (fixed)
336.17121
The two-planet MMR takes place when two values of pi are
non-zero. If three coefficients in this linear relation are non-zero,
it means that the system exhibits 3 -- body MMR. In the Kepler-11
system, the fundamental frequencies associated with ωi and Ωi are
much smaller than ni (these are the secular frequencies), hence the
right-hand sides of Eq. 3 are close to 0. This makes it possible to
skip the secular terms, as the first order approximation, and to iden-
tify the MMRs through approximate resonance conditions involv-
ing the mean motions only.
To identify the MMRs in the MEGNO maps, we apply a sim-
ple method described in (Guzzo 2005). In the (n j,nk) -- plane4, the
slope
α j,k ≡ dnk
dn j
of a particular resonance line determines the ratio of coefficients pi,
i.e.,
α j,k = − p j
pk
.
(4)
If α j,k = 0 then the MMR forms a horizontal line, planet k is in-
volved in the MMR, while planet j is not. If α−1
j,k = 0 then the
MMR forms a vertical line, planet j is involved in the resonance,
while planet k is not. In these cases, other planets may be involved
in this particular resonance. If α j,k is finite and non -- zero then both
considered planets are involved in the MMR. To identify this par-
ticular resonance, one has to compute slopes corresponding to this
resonance in all (ni,n j)-planes. It may be not possible, if the map
ranges do not cover the resonance band. If α j,k is non-zero and fi-
nite only for one pair of planets ( j, k), it means that 2 -- planet MMR
is present. It should be verified whether p jn j + pknk ≈ 0. Coef-
ficients p j, pk of the MMR condition can be computed from the
slopes α j,k. Similarly, the 3-body MMR takes place, if there exist
4 Let us note that Fig. 12 shows the (xi,x j)-planes but (ni,n j) may be easily
computed from these data.
Table 7. Three-planet resonances near the best -- fit model IIa. See Fig. 12.
label
pb
1
2
3
4
5
6
7
1∗
2∗
7
0
0
1
9
0
0
0
0
pc
−10
0
0
0
−13
0
0
5
6
pd
2
0
5
−10
0
6
6
−8
−14
pe
0
7
−5
11
4
−9
−16
−1
5
pf
0
−11
−3
0
0
0
11
0
0
pg
0
2
0
0
0
2
0
0
0
relatively prime integers i, j (cid:54)= i and k (cid:54)= i, j, such that αi, j,αi,k and
α j,k are all finite and non -- zero. The 3-body MMR may be iden-
tified by computing the slope coefficients in appropriate planes of
the mean motions. An identification of 4 -- body and N -- body MMRs
can be derived as well.
Using this simple MMR identification algorithm, we found
most significant MMRs close to solution IIa. The identified 3 -- body
MMRs were labeled at the panels, and listed in Tab. 7. The mean
motions of solution IIa permit a few low -- order 2 -- body MMRs in
the vicinity of this solution, e.g., 4nb − 5nc, 1nb − 3ne, 2nc − 5ne,
1nd − 2nf, 2ne − 3nf. There is no 2-planet MMR in the range of ni
implied by 1σ uncertainty. All straight bands with finite and non-
zero αi, j have at least two images in the planes constructed for other
planets. All features seen in the dynamical maps correspond then
to 3 -- and 4 -- body MMRs. Possibly, even more complex N-body
resonances may be found.
Labels in Fig. 12 corresponds to data in Tab. 7. Labels with
asterisks are written in open circles in the dynamical maps and de-
note MMRs in the neighborhood of solution IIa. Resonances la-
beled with numbers without asterisks and written in filled circles in
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
15
Table 8. Three-planet resonances near solution IIb
label
1
2
3
4
5
6
7
3∗
4∗
5∗
pb
pc
−10
7
0
0
0
0
1
0
−13
9
0
0
0
0
0
1
13 −17
0
0
pd
2
0
5
−10
0
6
6
−6
0
11
pe
0
7
−5
11
4
−9
−16
6
0
−21
pf
0
−11
−3
0
0
0
11
0
2
8
pg
0
2
0
0
0
2
0
0
0
0
argument of pericenter is shifted from the nominal value by ∆ωi,
the mean anomaly should be shifted by −∆ωi. For eccentric orbits
such a correction flows from the Ist Keplerian law.
The MMRs form even more sophisticated web in the (ωi,ω j) --
planes than in the mean motion planes. These dynamical maps re-
veal that the stability depends on the initial arguments of pericen-
ters. It is not obvious a priori, because eccentricities are very small.
The initial phases of the system are preserved across the maps,
and each point corresponds to the same Ti. Keeping in mind that
the photometric data spanning only ∼ 500 days imply weak con-
straints on angles ωi, we realize how is difficult to find a stable
initial conditions in the huge, 50-dimensional parameter space of
the system. Figure 16 illustrates MEGNO maps in the (ei,Ωi) -
planes calculated over T = 8000 yr. Each panel is for one planet.
An identification of particular MMRs is much more complex than
in the mean motion planes. It would require the frequency analysis
of these solutions. Still, regions of stable, quasi-periodic motions
usually form only small islands in the phase space. For the four
innermost planets, the regular motions are confined to only ∼ 5◦
range of Ωi around the nominal value and also to a small range of
eccentricities ∼ 0.01. For the two outermost planets f and planet g
the maps look like different. A zone of stable motions of planet g
extends towards large Ωg. It implies a large mutual inclination of
its orbit to the rest of the system. Small Ωg provoke unstable mo-
tions. The nominal solution is found at the very edge between the
regular and chaotic zone.
5.4 Stable solution of model IV
Let us recall that in transit model IV all Ωi = 0◦. Hence, the mutual
inclinations between all pairs of orbits are close to 0◦ but the system
remains non-coplanar because Ii is still different from 90◦ (hence,
the transits of all planets in this system can be detected by a ran-
domly located observer with a significant probability of ∼ 3.4%).
In this case we found several solutions with MEGNO converging
to 2 after T = 16000 yr. This indicates a possibility of quasi-regular
orbits. We chose one of such solutions. Its parameters are listed in
Tab. 9, and we compute dynamical maps in its vicinity. Figure 17
shows the results of this experiment in the mean motions planes.
The integration time is T = 8000 yr per pixel. The tested config-
urations is found in a narrow region of regular motions. The most
prominent dynamical feature visible in the maps is associated with
stable 3-planet MMR between planets b, c and d. We identified it
as (7,−10,2,0,0,0) MMR. It has been also found in dynamical
maps associated with solutions IIa and IIb as the MMR labeled as
"1". Due to altered parameters of the model, the unstable region of
Figure 13. Dynamical map of solution IIa but for the integration time
40000 yr per pixel. Mean motion resonances are labeled in accord with
Table 7.
the dynamical map are also present in the neighbourhood of solu-
tion IIb.
The MMRs (7ne−11nf +2ng) labeled as "2" and (5nc−8nd−
1ne) labeled as "1∗" are the most close to solution IIa. Solution IIa
passed the 16000 yr MEGNO test as stable solution. However, be-
cause it is close to two 3 -- body MMRs, and is found in a a dense
web of low -- order 3 -- body and 4-body MMRs, its long -- term sta-
bility cannot be guaranteed by this test. The integration time of
16000 yr corresponds to ∼ 50,000 orbital periods of the outermost
planet g. This time is usually too short to detect a chaotic nature of
the orbit when the 3 -- body resonances are present. Unfortunately,
the CPU-overhead does not permit to derive the dynamical map
with the integration time per single initial condition which should
be 10 -- 102 -- times longer. Indeed, a test run over T = 40,000 yr il-
lustrated in Fig. 13 reveals that solution IIa is chaotic and unstable.
Besides, almost the whole (xb,xc)-plane corresponds to chaotic mo-
tions with (cid:104)Y(cid:105) > 5.
We analyzed the second solution IIb in the same manner. The
MEGNO maps computed for 8000 yr are presented in Fig. 14. Also
these maps reveal a dense net of 3-body and 4-body resonances.
Some of these resonances may be identified as close to solutions IIa
as well as to IIb. They are labeled with the same numbers written
within filled circles, as in Fig. 12. We found also a few new MMRs,
labeled within open circles and labeled by "3", "4" and "5". The
remaining 4 -- body MMRs which are visible in this figure are not
labeled. All identified MMRs are listed in Tab. 8.
Similarly to configuration IIa, the integration over longer time
of T = 40000 yr, reveals that solution IIb is unstable. Almost the
whole plane of the dynamical map corresponds to (cid:104)Y(cid:105) > 5. The
MEGNO map is similar to Fig. 13, and is not shown here.
(ei,Ωi)
(ωi,ω j)
5.3 Dynamical maps in the (ωi,ω j)
(ωi,ω j) -- and (ei,Ωi)
(ei,Ωi)-planes.
Figure 15 illustrates the MEGNO maps in planes of the arguments
of pericenters. The MEGNO is calculated over T = 8000 yr. The
(ωi,ω j) -- maps are constructed a bit differently than the mean mo-
tions dynamical maps. Because we intent to analyse configurations
coherent with the observations, ωi can be freely varied over the
grid, if also the mean anomalies are modified to preserve the time
of the first transit, Ti. For instance, for a circular orbit, when the
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
16
C. Migaszewski, M. Słonina and K. Go´zdziewski
Figure 14. Symplectic MEGNO maps in the (xi,x j) -- plane for solution IIb. Blue color means (cid:104)Y(cid:105) ≈ 2 (stable configuration), while yellow is for (cid:104)Y(cid:105) (cid:38) 5 and
unstable systems. Each panel is for different pair of planets labeled in the bottom-right corner of each particular panel. The resolution is 512× 512 pixels, total
integration time is 8000 yr per pixel, SABA4 integrator step-size is 0.5 days.
Table 9. Orbital parameters of solution IVa. Mass of the star is 0.95m(cid:12). Osculating elements of Poincar´e are given at the epoch JD 2455964.51128.
parameter/planet
b
c
d
e
f
g
m [m⊕]
a [au]
e
I [deg]
ω [deg]
M [deg]
2.359
0.091113
0.04423
89.141
20.651
3.386
0.106505
0.01719
91.215
55.728
178.88174
209.60077
5.630
0.154243
0.00633
89.332
140.753
40.79259
10.841
0.193940
0.00258
88.837
236.761
318.51831
7.524
0.249511
0.01073
89.394
355.845
91.57569
25.161
0.463991
0 (fixed)
89.770
0 (fixed)
336.26502
this resonance is visible in these maps. The solid black lines plotted
across panels shown in the top row of Fig. 17 have the slope equal to
7/10, −7/2 and 5, from the left to the right, respectively. Other 3-
body resonances visible in the dynamical maps form a dense web,
which may be better seen in the (xd,xe) -- and (xd,xf) -- planes dis-
played in the bottom panels.
To examine the stability of this solution over longer time scale,
we calculated a single dynamical map in the (xb,xc) -- plane for
much longer integration time T = 40000 yr. It is shown in Fig. 18.
This solution is located in a tiny region of stable motions. In this
case, the 3 -- body MMRs has a protective role for the system, saving
it from a disruption. Actually, this solution is chaotic. To demon-
strate this, we computed the critical argument of the 3 -- body MMR
and we plot over two intervals of time, at the beginning of the 1 Myr
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
17
Figure 15. Dynamical maps in the (ωi,ω j) -- plane for solution IIa. The MEGNO range is [1,5] and colour -- coded: blue is for stable solutions, yellow is for
unstable systems. Each map has elements of a given pair of planets varied, these planets are labeled in the bottom-right corner. The arguments of pericenter
are expressed in degrees. The nominal solution is marked with the asterisk.
integration period (the left panel of Fig. 19) and at the end of this
period (the right panel of Fig. 19). The critical arguments exhibit
librations alternated with circulations. Such behavior of the critical
argument indicates a crossing of the separatrix of the resonance,
and chaotic dynamics. In such a case, the configuration may be
geometrically stable over very long time but it may be suddenly
disrupted due to a slow diffusion along the resonance.
5.5 The Arnold web structure in the phase space
The results of experiments described in the previous sections may
be interpreted at the ground of the dynamical systems theory. The
dynamical stability of planetary orbits in systems with strong per-
turbations is influenced by even small errors and the resulting dif-
fusion due to resonances overlapping. This dynamical phenomenon
has been investigated in the Outer Solar system. Murray & Hol-
man (2001) identified the chaos among the Jovian planets as result-
ing from the overlap of the components of 3 -- body MMRs among
Jupiter, Saturn, and Uranus. In spite of short Lyapunov time (107
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
years), they found the dynamical lifetime of Uranus ∼ 1018 years.
In this way, the analytic theory of Murray & Holman (2001) ex-
plained an apparent paradox of long -- term stable system which is
actually chaotic.
The structure of the 3 -- body and 4 -- body MMRs in the Outer
Solar system was further investigated numerically by Guzzo (2005,
2006). Very recently, it was also studied by Quillen (2011) for
strongly interacting extrasolar systems. Through the dynamical
maps technique, Guzzo (2005) found that if the chaotic motions
appear in a regular net, they may be practically stable over very
long times. Such a state of chaotic system is called the Nekhoro-
shev regime. On contrary, if the chaotic resonances do not consti-
tute a regular web, and minority of orbits form a global chaotic
zone, the stability of the system is influenced by strong chaotic
diffusion. Such regime is related to the resonance overlap, and is
called the Chirikov regime of the dynamics (Froeschl´e et al. 2000;
Guzzo 2005).
These results may be applied to interpret the dynamical maps
of the Kepler-11 system. The dense net of the multiple-body MMRs
form the Arnold web in the neighborhood of the best fit model
18
C. Migaszewski, M. Słonina and K. Go´zdziewski
Figure 16. Dynamical maps in the (ei,Ω j)-plane for solution IIa. The MEGNO range [1,5] is colour coded: blue means stable solutions and yellow is for
unstable configurations. Each map is constructed for a pair of planets labeled in the bottom-right corner. The longitudes of ascending nodes are expressed in
degrees. The nominal solution is marked with an asterisk. See Tab. 5 for the orbital elements of this best -- fit model.
Figure 17. Dynamical maps of solution IVa in the mean motions planes. The colours and symbols are the same as in Fig. 12.
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
19
Figure 19. Temporal evolution of the critical argument of the 3-body MMR for the initial condition IVa. The left panel is for the beginning of the integration
period, and the right panel is for the end of the integration period spanning 1 Myr.
of long -- term integrations of the obtained sets of initial conditions
indicate that the system resides at the edge of dynamical stability.
These conclusions might be changed, if more data are gathered and
analyzed.
6 CONCLUSIONS
In this paper, we derived an improved method for the dynamical
analysis of photometric light curves of stars with multiple transit-
ing planets. Its main purpose is to determine planetary masses, as
well as a number of indirect parameters affecting the dynamical sta-
bility of the system. This algorithm improved the well known TTV
algorithm. The crucial point of this method is to model the whole
photometric curve directly with the help of an efficient symplectic
N-body integration. Such the direct approach make is possible to
account for multiple transits, as well for the transit depths and their
widths. This in turn makes it possible to impose dynamical con-
straints on parameters which cannot be determined in terms of the
TTV, like the longitude of nodes and mutual inclinations of orbits.
With the help of this new method, we re-analyzed available
photometric data for Kepler-11. The direct algorithm imposes con-
straints on the mass of the outermost planet g and help us to de-
termine the mutual inclinations between orbits of planets b and
planet c as well as between the (d-e) pair with a good accuracy of
2◦. These results extend analysis performed in the discovery paper
(Lissauer et al. 2011). Overall, conclusions in this paper and in our
work coincide very well, in spite of quite a different transit mod-
els, optimisation algorithms and uncertainties estimation methods
applied.
Thanks to the in -- depth analysis of the Kepler-11 light-curves,
we investigated a possible chemical composition of the planets de-
tected in this intriguing system. The most curious finding is a clear
anti-correlation of the mean densities of the planets with their mean
distances from the star. The inner planets exhibit larger abundance
of heavy elements than the outer companions. Because all eccen-
tricities as well as the mutual inclinations of stable systems remains
small, the system unlikely suffered violent scattering processes in
the past. It follows that the ordering of planets have been preserved
since their formation. A dynamical relaxation should imply large
ei and ∆Ii, j (see, e.g., Chatterjee et al. 2008; Adams & Laugh-
lin 2003). These factors indicate that the primordial protoplanetary
disk might have a significant gradient of metallicity and the present
Kepler-11 system is a real fossil of this disk. If this suggestion is
confirmed, it can constrain the planet formation theories, in partic-
ular concerning multiple systems of super -- Earth planets.
This conclusion is reinforced by the dynamical analysis of the
system. We found, in accord with the discovery paper, that the sys-
tem is basically free from 2-planet MMRs. However, its global dy-
Figure 18. Dynamical map of solution IVa in the mean motion plane of
planet b and planet c but for longer integration time of T = 40,000 yr. See
Fig. 17 for an explanation of the symbols and colour coding.
configurations. Our experiments reveal, that this system may un-
dergo the Chirikov regime rather than long-term stable Nekhoro-
shev regime. We have no strong proof of this phenomenon, be-
cause it would require non-trivial and intensive computations of
the chaotic diffusion. Conclusions regarding the real state of the
Kepler-11 system require more precise determination of the initial
conditions than obtained in this work.
We also note that solutions IIa, IIb investigated in detail are
found at the very border of the chaotic and regular zones. This
can be well seen in the (ωi,ω j) -- and (ei,Ωi)-planes. A change of
these angles constrained by the best -- fit errors, could "move" the
system into larger zones of stable motions. Because the MEGNO
quantifies the stability of the system as a whole, such a change
would imply a shift of the initial condition in all parameter maps
(Go´zdziewski & Maciejewski 2001). The altered initial conditions
would be also more "distant" from the unstable strips of 3 -- body
and 4 -- body MMRs in the mean motions planes. The dynamical
maps would be more similar to those computed for solution IV re-
vealing most extended zones of stable motions. Overall, the dy-
namical state of the system is very fragile and depends on subtle
changes of the initial conditions.
Still most of the best-fit model configurations obtained with
the bootstrap algorithm, which appear as regular over the short-
term dynamics time -- scale are self -- disrupting. This behavior is
similar to unstable evolution in 3 -- body MMRs observed in the
HD 37124 system (Go´zdziewski et al. 2008). In this sense the dy-
namical state of the Kepler-11 system is still puzzling. The results
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
20
C. Migaszewski, M. Słonina and K. Go´zdziewski
namics is governed by 3-body and 4 -- body MMRs. Overlapping
of these resonances near the best-fit solutions imply an extended
zone of dynamical chaos and very unstable configurations. Partic-
ular multi-body resonances may stabilize the system. We identified
such a 3 -- body resonance. However, the observation window span-
ning only ∼ 500 days does not make it possible to pick up this
MMR as the only possible. Besides, the MMRs form the structure
of the Arnold web characteristic for the Chirikov regime. In this
dynamical state, the phase-space orbits are strongly unstable due to
overlapping of the MMRs. In such a case, the chaotic diffusion in
the actions (semi -- major axes) space is significant and easily leads
to strong geometric changes of orbits. Actually, our numerical ex-
periments indicate this in the Kepler-11 system. It remains an open
question though, how such an apparently ordered configuration of
planets could survive the formation phase in extremely complex
and fragile dynamical environment.
The discovery team argue that the system is non-resonant.
This factor would prevent a scenario of trapping the planets into
MMRs during the inward migration at the early stages of the evo-
lution. As we showed here, the system is in fact extremely resonant,
but in quite a different sense. Combining this fact with small val-
ues of the eccentricities and anti -- correlation of the chemical com-
position with the distance to the star, we may conclude quite an
opposite: the migration/trapping scenario could be the only way of
preserving the primary architecture in the present form. However,
these suggestions might be verified only if more photometric data
are gathered and are available.
Most likely, many other Kepler -- discovered multiple extraso-
lar systems with transiting planets exhibit qualitatively similar be-
haviours to that one we found in the Kepler-11. Hence, the ap-
proach in this paper is general and may be applied in the studies
of other compact systems composed of low-massive, super -- Earth
or Neptune-like planets.
ACKNOWLEDGMENTS
We would like to thank the anonymous referee for a review and
comments that improved the text. This work is supported by
the Polish Ministry of Science and Higher Education grant No.
N/N203/402739. CM is a recipient of the Foundation for Polish
Science Fellowship (programme START, editions 2010 and 2011).
This research was carried out with the support of the "HPC Infras-
tructure for Grand Challenges of Science and Engineering" project
(POWIEW), co-financed by the European Regional Development
Fund under the Innovative Economy Operational Programme.
REFERENCES
Adams F. C., Laughlin G., 2003, Icarus, 163, 290
Agol E., Steffen J., Sari R., Clarkson W., 2005, MNRAS, 359, 567
Baluev R. V., 2009, MNRAS, 393, 969
Barnes R., Go´zdziewski K., Raymond S. N., 2008, ApJL, 680,
L57
Batalha N. M., Borucki W. J., Bryson S. T., Buchhave L. A., Cald-
well D. A., Christensen-Dalsgaard J., Ciardi D., Dunham E. W.,
Fressin F., Gautier III T. N., Gilliland R. L., Haas M. R., Howell
S. B., Jenkins J. M., Kjeldsen H., Koch D. G., 2011, ApJ, 729,
27
Batalha et al. N. M., 2012, ArXiv 1202.5852
Borucki W. J., Koch D., Basri G., Batalha N., Brown T., Cald-
well D., Caldwell J., Christensen-Dalsgaard J., Cochran W. D.,
DeVore E., Dunham E. W., 2010, Science, 327, 977
Borucki et al. W. J., 2011, ApJ, 736, 19
Caldwell D. A., Kolodziejczak J. J., Van Cleve J. E., Jenkins J. M.,
Gazis P. R., Argabright V. S., Bachtell E. E., Dunham E. W.,
Geary J. C., Gilliland R. L., Chandrasekaran H., Li J., 2010,
ApJL, 713, L92
Carter J. A., Agol E., Chaplin W. J., Basu S., Bedding T. R., Buch-
have L. A., Christensen-Dalsgaard J., Deck K. M., Elsworth Y.,
Fabrycky D. C., Ford E. B., Fortney J. J., Hale S. J., Handberg
R., Hekker S., Holman M. J., Huber D., 2012, ArXiv e-prints
Carter J. A., Fabrycky D. C., Ragozzine D., Holman M. J., Quinn
S. N., Latham D. W., Buchhave L. A., Van Cleve J., Cochran
W. D., Cote M. T., Endl M., Ford E. B., Haas M. R., Jenkins
J. M., 2011, Science, 331, 562
Charbonneau P., 1995, ApJS, 101, 309
Chatterjee S., Ford E. B., Matsumura S., Rasio F. A., 2008, ApJ,
686, 580
Cincotta P. M., Giordano C. M., Sim´o C., 2003, Physica D Non-
linear Phenomena, 182, 151
Cincotta P. M., Sim´o C., 2000, A&A Supl., 147, 205
Cochran W. D., Fabrycky D. C., Torres G., Fressin F., D´esert
J.-M., Ragozzine D., Sasselov D., Fortney J. J., Rowe J. F.,
Brugamyer E. J., Bryson S. T., Carter J. A., Ciardi D. R. e. a.,
2011, ApJS, 197, 7
Deb K., 2004, IPSJ Transactions on Mathematical Modeling and
Its Applications, 45, 1
Fabrycky D. C., Ford E. B., Steffen J. H., Rowe J. F., Carter J. A.,
Moorhead A. V., Batalha N. M., Borucki W. J., Bryson S., Buch-
have L. A., Christiansen J. L., Ciardi D. R., Cochran e. a., 2012,
ArXiv e-prints
Ford E. B., Fabrycky D. C., Steffen J. H., Carter J. A., Fressin F.,
Holman M. J., Lissauer J. J., Moorhead A. V., Morehead R. C.,
Ragozzine D., Rowe J. F., Welsh W. F., Allen C., Batalha N. M.,
Borucki W. J., Bryson S. T., Buchhave L. A. e. a., 2012, ArXiv
e-prints
Fressin F., Torres G., D´esert J.-M., Charbonneau D., Batalha
N. M., Fortney J. J., Rowe J. F., Allen C., Borucki W. J., Brown
T. M., Bryson S. T., Ciardi D. R., Cochran W. D., Deming D.,
Dunham E. W., Fabrycky D. C., Gautier III T. N., Gilliland R. L.,
2011, ApJS, 197, 5
Fressin F., Torres G., Rowe J. F., Charbonneau D., Rogers L. A.,
Ballard S., Batalha N. M., Borucki W. J., Bryson S. T., Buchhave
L. A., Ciardi D. R., Desert J.-M., Dressing C. D., Fabrycky D. C.,
2011, ArXiv e-prints
Froeschl´e C., Guzzo M., Lega E., 2000, Science, 289, 2108
Gautier III T. N., Charbonneau D., Rowe J. F., Marcy G. W., Isaac-
son H., Torres G., Fressin F., Rogers L. A., D´esert J.-M., Buch-
have L. A., Latham D. W., Quinn S. N., Ciardi D. R., Fabrycky
D. C., Ford E. B., Gilliland R. L., Walkowicz L. M., 2011, ArXiv
e-prints
Go´zdziewski K., 2003, A&A, 398, 315
Go´zdziewski K., Breiter S., Borczyk W., 2008, MNRAS, 383, 989
Go´zdziewski K., Maciejewski A. J., 2001, ApJL, 563, L81
Go´zdziewski K., Migaszewski C., Musieli´nski A., 2008, in Y.-
S. Sun, S. Ferraz-Mello, & J.-L. Zhou ed., IAU Symposium
Vol. 249 of IAU Symposium, Stability constraints in modeling
of multi-planet extrasolar systems. pp 447 -- 460
Guillot T., 1999, Science, 296, 72
Guzzo M., 2005, Icarus, 174, 273
Guzzo M., 2006, Icarus, 181, 475
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
A dynamical analysis of the Kepler-11 planetary system
21
Helled R., Anderson J. D., Podolak M., Schubert G., 2011, ApJ,
726, 15
Holman M. J., Fabrycky D. C., Ragozzine D., Ford E. B., Stef-
fen J. H., Welsh W. F., Lissauer J. J., Latham D. W., Marcy
G. W., Walkowicz L. M., Batalha N. M., Jenkins J. M., Rowe
J. F., Cochran W. D., Fressin F., Torres G., 2010, Science, 330,
51
Holman M. J., Murray N. W., 2005, Science, 307, 1288
Howarth I. D., 2011, MNRAS, 418, 1165
Jenkins J. M., Caldwell D. A., Chandrasekaran H., Twicken J. D.,
Bryson S. T., Quintana E. V., Clarke B. D., Li J., Allen C., Tenen-
baum P., Wu H., Klaus T. C., Middour C. K., Cote M. T., Mc-
Cauliff S., Girouard F. R., 2010, ApJL, 713, L87
Koch D. G., Borucki W. J., Basri G., Batalha N. M., Brown T. M.,
Caldwell D., Christensen-Dalsgaard J., Cochran W. D., DeVore
E., Dunham E. W., Gautier III T. N., Geary J. C., Gilliland R. L.,
Gould A., Jenkins J., 2010, ApJL, 713, L79
Laskar J., Robutel P., 2001, Celestial Mechanics and Dynamical
Astronomy, 80, 39
Lissauer J. J., Fabrycky D. C., Ford E. B., Borucki W. J., Fressin
F., Marcy G. W., Orosz J. A., Rowe J. F. ., 2011, Nature, 470, 53
Lissauer J. J., Marcy G. W., Rowe J. F., Bryson S. T., Adams E.,
Buchhave L. A., Ciardi D. R., Cochran W. D., Fabrycky D. C.,
Ford E. B., Fressin F., Geary J., Gilliland R. L., Holman M. J.,
Howell S. B. e. a., 2012, ArXiv e-prints
Lynden-Bell D., O'Dwyer J. P., 2001, ArXiv Astrophysics e-
prints
Lynden-Bell D., Tout C. A., 2001, ApJ, 558, 1
Mandel K., Agol E., 2002, ApJ, 580, L171
Mikkola S., Innanen K., 1999, Celestial Mechanics and Dynami-
cal Astronomy, 74, 59
Morbidelli A., 2002, Modern celestial mechanics : aspects of solar
system dynamics
Murray N., Holman M., 2001, Nature, 410, 773
Nesvorn´y D., Kipping D. M., Buchhave L. A., Bakos G. ´A., Hart-
man J., Schmitt A. R., 2012, Science, 336, 1133
P´al A., 2011, MNRAS, p. 2123
Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P.,
1992, Numerical recipes in C. The art of scientific computing
Quillen A. C., 2011, MNRAS, 418, 1043
Rogers L. A., Bodenheimer P., Lissauer J. J., Seager S., 2011,
ApJ, 738, 59
Slonina M., Gozdziewski K., Migaszewski C., 2012, ArXiv e-
prints
Steffen J. H., Fabrycky D. C., Ford E. B., Carter J. A., D´esert J.-
M., Fressin F., Holman M. J., Lissauer J. J., Moorhead A. V.,
Rowe J. F., Ragozzine D., Welsh W. F., Batalha N. M., Borucki
W. J., Buchhave L. A., Bryson S. e. a., 2012, MNRAS, 421, 2342
Wisdom J., Holman M., 1991, AJ, 102, 1528
c(cid:13) 2012 RAS, MNRAS 000, 1 -- 21
|
1304.4204 | 2 | 1304 | 2015-09-11T18:18:25 | A Distribution of Large Particles in the Coma of Comet 103P/Hartley 2 | [
"astro-ph.EP"
] | The coma of comet 103P/Hartley 2 has a significant population of large particles observed as point sources in images taken by the Deep Impact spacecraft. We measure their spatial and flux distributions, and attempt to constrain their composition. The flux distribution of these particles implies a very steep size distribution with power-law slopes ranging from -6.6 to -4.7. The radii of the particles extend up to 20 cm, and perhaps up to 2 m, but their exact sizes depend on their unknown light scattering properties. We consider two cases: bright icy material, and dark dusty material. The icy case better describes the particles if water sublimation from the particles causes a significant rocket force, which we propose as the best method to account for the observed spatial distribution. Solar radiation is a plausible alternative, but only if the particles are very low density aggregates. If we treat the particles as mini-nuclei, we estimate they account for <16-80% of the comet's total water production rate (within 20.6 km). Dark dusty particles, however, are not favored based on mass arguments. The water production rate from bright icy particles is constrained with an upper limit of 0.1 to 0.5% of the total water production rate of the comet. If indeed icy with a high albedo, these particles do not appear to account for the comet's large water production rate. production rate.
Erratum: We have corrected the radii and masses of the large particles of comet 103P/Hartley 2 and present revised conclusions in the attached erratum. | astro-ph.EP | astro-ph |
A distribution of large particles in the coma of Comet
103P/Hartley 2
Michael S. Kelleya,∗, Don J. Lindlerb, Dennis Bodewitsa, Michael F. A'Hearna, Carey M.
Lissec, Ludmilla Kolokolovaa, Jochen Kisseld,1, Brendan Hermalyne
aDepartment of Astronomy, University of Maryland, College Park, MD 20742-2421, USA
bSigma Space Corporation, 4600 Forbes Boulevard, Lanham, MD 20706, USA
cJohns Hopkins University -- Applied Physics Laboratory, 11100 Johns Hopkins Road, Laurel, MD 20723,
dMax-Planck-Institut fur Sonnensystemforschung, Max-Planck-Str. 2, 37191 Katlenburg-Lindau, Germany
eNASA Astrobiology Institute, Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive,
Honolulu, HI 96822, USA
USA
Abstract
The coma of Comet 103P/Hartley 2 has a significant population of large particles observed
as point sources in images taken by the Deep Impact spacecraft. We measure their spatial
and flux distributions, and attempt to constrain their composition. The flux distribution of
these particles implies a very steep size distribution with power-law slopes ranging from −6.6
to −4.7. The radii of the particles extend up to 20 cm, and perhaps up to 2 m, but their
exact sizes depend on their unknown light scattering properties. We consider two cases:
bright icy material, and dark dusty material. The icy case better describes the particles
if water sublimation from the particles causes a significant rocket force, which we propose
as the best method to account for the observed spatial distribution. Solar radiation is a
plausible alternative, but only if the particles are very low density aggregates. If we treat
the particles as mini-nuclei, we estimate they account for < 16 − 80% of the comet's total
water production rate (within 20.6 km). Dark dusty particles, however, are not favored based
on mass arguments. The water production rate from bright icy particles is constrained with
an upper limit of 0.1 to 0.5% of the total water production rate of the comet.
If indeed
icy with a high albedo, these particles do not appear to account for the comet's large water
production rate.
Erratum: We have corrected the radii and masses of the large particles of comet
103P/Hartley 2 and present revised conclusions in the attached erratum.
∗Corresponding author
Email addresses: [email protected] (Michael S. Kelley), [email protected] (Don J.
Lindler), [email protected] (Dennis Bodewits), [email protected] (Michael F. A'Hearn),
[email protected] (Carey M. Lisse), [email protected] (Ludmilla Kolokolova),
[email protected] (Brendan Hermalyn)
1retired
Kelley et al. 2013, Icarus 222, 634-652
1. Introduction
Comet 103P/Hartley 2 (hereafter, 103P or Hartley 2) is a hyperactive comet. The ac-
tivities of most comets are consistent with surfaces that are over 90% inert, i.e., activity is
restricted to a few localized sources. In contrast, the gas production rates of the hyperactive
comets suggest activity over ∼ 100% of their surfaces or more. Comet Hartley 2's peak
water production rate during the 1997 perihelion passage was QH2O ≈ 3 × 1028 molec s−1
(Combi et al. 2011). The sublimation rate of an isothermal nucleus with a 4% Bond albedo
at 1.03 AU (the 1997 perihelion distance of Hartley 2) is 2.9 × 1017 molec cm2 s−1 (Cowan
and A'Hearn 1979). Thus, approximately 10 km2 of surface area would be required to match
the measured water production in 1997. However, constraints on the size of the nucleus by
Groussin et al. (2004) and Lisse et al. (2009) suggested the total surface area was near 4 --
6 km2, yielding an active fraction (area based on QH2O/area of the nucleus) near 1.7 -- 2.5. In
order to account for the high active fraction, Lisse et al. (2009) proposed that the coma con-
tained a population of icy grains, which increased the surface area available for sublimation,
and therefore allowed for a relatively high water production rate.
The Deep Impact spacecraft flew by Comet 103P on 4 November 2010. The minimum
flyby distance was 694 km, occurring at 13:59:47.31 UTC with a relative speed of 12.3 km s−1
(A'Hearn et al. 2011). Rather than just being the fifth comet nucleus to be imaged up close
by a spacecraft, Comet Hartley 2 might also be considered an archetype of the hyperactive
comets. The flyby images verified the small surface area of the nucleus (5.24 km2; Thomas
et al. this issue), and Deep Impact's IR spectrometer revealed a coma of icy grains (A'Hearn
et al. 2011). Knight and Schleicher (this issue) discovered tail of OH and NH, evident on
104 km scales, which they conclude are derived from small grains of ice that were accelerated
down the tail before completely sublimating. Bonev et al. (this issue) observed enhanced
heating of the water gas 20 -- 150 km down the tail, which they also attribute to sublimating
ice grains in the tail. The icy coma hypothesis of Lisse et al. (2009) seems to be qualitatively
valid. What remains is a quantitative verification that the ice grains in the coma are a
significant source of water gas.
A'Hearn et al. (2011) also discovered bright point sources distributed around the nucleus
of Hartley 2 in visible wavelength images. The fluxes of these sources are consistent with a
population of objects centimeter-sized or larger. A'Hearn et al. (2011) could not determine if
these large particles were icy or refractory. In the following paper, we will attempt to discern
the nature of these point sources and if they might be the cause of Hartley 2's hyperactivity.
The point sources around Hartley 2 are found in visible wavelength images taken with
both the High Resolution Instrument (HRI) and the Medium Resolution Instrument (MRI).
Their spatial distributions and densities vary, but they are not background stars; celestial
sources are significantly streaked in images taken while the spacecraft-comet distance, ∆,
was less than 104 km. That the HRI detected these point sources is significant. The HRI
visible camera (HRIVIS) is not in optimal focus, and has an accordingly large and doughnut-
shaped point spread function (PSF) (Lindler et al. 2007). Thus, unresolved sources are easily
discriminated from cosmic ray impacts and hot pixels. None of the point sources appear to
be resolved, therefore they must be smaller than the FWHM of the HRIVIS deconvolved
PSF (approximately 3 m at ∆ = 700 km).
In addition to point sources, streaks of various lengths are present in many of the images.
2
The lengths of the streaks are correlated with exposure time (longer exposures produce longer
streaks), and inversely proportional to instrument pixel scale (streaks in HRIVIS images are
longer than in MRI). The streaks are parallel to each other, and the orientation is consistent
with the spacecraft's motion as it tracks 103P's nucleus. Stars and other distant objects
produce streaks of uniform length and orientation, but we find a variety of streak lengths
in a single image, down to a few pixels long, which indicates they are spatially close to the
nucleus and not celestial sources.
Altogether, the above observations indicate that the nucleus of Comet 103P is surrounded
by a coma of thousands of large particles. We will use the term "particle" to refer to
the observed point sources in this paper. A'Hearn et al. (2011) demonstrated that these
particles are centimeter-sized or larger. Large particles are not unique to this comet, but
these observations are unique to large particles. These are the first images of sub-meter
particles from a comet seen as individual objects. Prior images in visible/infrared light
(e.g., comet dust trails; Sykes and Walker 1992, Ishiguro et al. 2002, Reach et al. 2007)
and with radar (Harmon et al. 2004) probed collections of many large particles. The only
other observations of large individual particles from comets that we can think of are those
of meteors (from meteor streams associated with particular comets) and milligram particle
impacts on spacecraft. Although the latter examples may only be of order 0.1 -- 1 mm in size.
In the following paper, we present our methods for detecting and measuring the photo-
metric properties of the large particles in the coma of Comet 103P. We will convert their
fluxes into sizes, discuss their spatial distribution, attempt to constrain their composition,
and compare them to large particles observed in radar observations of the comet. Assuming
they are icy, we also constrain their contribution to the comet's total water production rate.
2. Observations
Images of Comet 103p Were taken with Deep Impact's Medium Resolution Instrument
and High Resolution Instrument CCD cameras. Both cameras have 1024× 1024 pixel arrays;
the HRIVIS pixel scale is 0.413(cid:48)(cid:48) pixel−1 (2 µrad), and the MRI pixel scale is 2.06(cid:48)(cid:48) pixel−1
(10 µrad). The instruments are described by Hampton et al. (2005), and their calibration
by Klaasen et al. (2008, in prep.). The MRI and HRIVIS data are available in the NASA
Planetary Data System (PDS) archive (McLaughlin et al. 2011a,b). Images taken with the
MRI are labeled with the prefix mv, and for the HRIVIS the prefix is hv. We keep this
convention throughout the paper. We used a pre-PDS data set, but the only significant
difference is the absolute flux calibration constant, which we account for in our photometry.
Optimal focus for the HRIVIS instrument occurs about 6 mm before the focal plane.
For this instrument, it is often beneficial to work with spatially restored (i.e., deconvolved)
images. We follow the method of Lindler et al. (2007), using the HRIVIS PSF from the
EPOXI mission (Barry et al. 2010), to restore raw HRIVIS images to near diffraction-limited
resolution. The HRIVIS images are deconvolved (Fig. 6b) with the Richardson-Lucy (R-L)
method modified to handle non-Poisson CCD readout noise (Snyder et al. 1993, Lindler et al.
2007). Note that R-L restoration methods conserve flux.
The large particles are easily seen in MRI and HRIVIS images at ∆ (cid:46) 104 km. Outside
of this range, the particles become increasingly faint, overwhelmed by the diffuse coma's
surface brightness, and confused with stars, which are unstreaked point sources when the
3
spacecraft-comet distance is large.
In a manual search, the earliest image in which we
have identified particles is HRIVIS image hv5000096, taken 22,400 km from the nucleus
(CA-30 min, 1500.5 ms exposure time). The first MRI image in which we have positively
identified particles is mv5002025 (8580 km, CA-11 min, 500.5 ms).
Most images at ∆ < 104 km use the CLEAR1 filter. For sources with solar-like spectra,
the CLEAR1 filters have effective wavelengths of λe = 0.625 µm for HRIVIS, and 0.610 µm for
MRI, and a FWHM of ≈ 0.5 µm. The absolute calibration uncertainties for the CLEAR1
filters are 5% for HRIVIS, and 10% for MRI (Klaasen et al. 2008). Large particles have
not yet been identified in the HRI infrared spectrometer (HRIIR) scans (Protopapa et al.
2011), but they may be discovered in the future when we can predict the location of the
brightest particles in IR data using their MRI and HRIVIS derived 3D positions and velocities
(Hermalyn et al. this issue).
A summary of all images used in this paper is presented in Table 1. Note that the HRIVIS
images used in this paper are only a small sub-set of the HRIVIS particle observations. We
have limited our investigation to the closest-approach images because of the great amount
of time that must be taken to analyze the HRIVIS data. We also do not use any of the
50.5 ms MRI images in the PDS archive. These images were taken along with the 500.5 ms
images, but few particles are found in these images due to the decreased sensitivity and great
distances from the nucleus.
3. Detection, photometry, and completeness
3.1. MRI images
We automatically detect and measure all point sources throughout all MRI images taken
within ∆ < 5200 km of the comet nucleus using DAOPHOT as included in the IRAF software
package (Tody 1993). In summary, we: (1) estimate the PSF of the instrument and verify
that it accurately measures particle fluxes; (2) remove the diffuse coma from the images;
(3) mask the nucleus and bright jets; (4) additionally mask stars in images at large ∆;
(5) measure the completeness of our particle census (i.e., detection efficiency); (6) estimate
the photometric uncertainties, and correct for any biases; and, (7) clean bad PSF fits and
crowded regions from our photometry lists. Below, we provide details on our methods.
Particles were detected by searching for point sources (FWHM = 1.5 pixels) with a
peak amplitude greater than 5σ above the local background: 1σ per pixel ≈ 1.1 × 10−3,
0.52× 10−3, and 0.13× 10−3 W m−2 µm−1 sr−1 for 41, 121, and 501 ms frames, respectively.
Twenty-five isolated and bright point sources were selected by hand from image mv5004056
to estimate the MRI PSF. The sources were fit by DAOPHOT with a variety of functions,
and the best-fit PSF used a Moffat function (β parameter of 1.5; Moffat 1969) with a look-
up table of empirical residuals. We compared DAOPHOT-derived fluxes of bright isolated
particles to those measured with aperture photometry. The values agreed, with a mean
error of ∆Fλ/Fλ = −1.9%, where ∆Fλ is the difference between the DAOPHOT flux and
the aperture flux (Fλ,DAOPHOT − Fλ,aper).
DAOPHOT iteratively estimates the background as it fits each point source and groups of
point sources, but we found -- especially for detections closer to the nucleus -- that subtracting
an initial estimate of the background improved DAOPHOT's ability to find and measure
4
sources. Furthermore, the MRI CCD has an instrumental background that varies row-
by-row at the 1 -- 4 DN level (the MRI and HRIVIS calibration constants for CLEAR1 are
3.527× 10−5 and 1.209× 10−4 W m−2 µm−1 sr−1 DN−1 s, respectively). The EPOXI pipeline
includes a routine to detect and remove this background, but it is not executed on images
dominated by coma, such as those in our analysis. We derive our background estimate for
each image using two median filters (i.e., low-pass filters). We first apply a 3× 25 pixel (row
× column) filter that defines the background stripes, as well as much of the coma. A second
25 × 3 pixel filter defines most of the residual coma, especially where it is extended along
image columns. A median-filter subtracted image is presented in Fig. 1b.
Bright jets and the nucleus complicate the PSF fitting. Therefore, we mask these regions
on the image. The mask is derived from a morphological gradient with an 11 × 11 pixel
uniform structuring element, which effectively generates an image of peak-to-peak values in
a moving 11 × 11 pixel box (Fig. 1c). We threshold the gradient image at a value of 0.1
W m−2 µm−1 sr−1, then dilate the mask by 30 pixels to fill small gaps between features, and
cover some of the nearby jets (Fig. 1d).
DAOPHOT will occasionally fit streaks in the MRI images with PSFs. Most of these
fits have poor residuals, and are removed before the final analysis (described below). But
we found several streaks persisting into the final analysis in images taken at large distances
from the comet (∆ (cid:38) 5000 km). These streaks were stars and not particles. DAOPHOT
seemed more likely to fit streaks in these images because the stars are much brighter than
the particles, and are only streaked several pixels or less. We generated an additional mask
by searching for elongated sets of pixels above the background in order to remove streaked
celestial objects from these images before processing with DAOPHOT. At closer comet-
spacecraft distances, stars are streaked over a greater number of pixels (up to a few hundred
pixels at closest approach), and are easily rejected without the mask.
We verified the completeness (i.e., detection efficiency) of our method by inserting artifi-
cial point sources into each image, re-executing our scripts on the new images, and examining
the output to determine if the artificial sources were detected. The completeness is then the
fraction of detected artificial sources per flux bin. We inserted 1% of the number of detected
point sources (with a minimum of 10 new particles) uniformly over the image but avoiding
the nucleus. The fluxes were picked from a distribution uniform in log-space, based on the
minimum and maximum fluxes measured in the image. The process is repeated, each time
starting with the original image, until 3000 artificial particles have been tested. Examples
of our MRI completeness test results are shown in Fig. 2. The completeness is directly cor-
related with exposure time; in general, it is easier to find and measure fainter particles in
images with longer exposures. In our MRI population studies, we will only consider particles
brighter than the 80% completeness level for each image (listed in online Table 2 as Fλ,min).
The completeness tests can be used to derive the photometric uncertainties. As an
example, we plot the input flux versus the DAOPHOT flux error for all artificial particles
added to image mv5004046 (Fig 3). In this image, the mean flux error ∆Fλ/Fλ = −8%,
and the mean error for all images ranges from −10% to −2%. However, the distribution is
not Gaussian as there is a significant tail at large negative errors. If we instead take the
resistant mean (clipping at 3σ), we find the mean ∆Fλ/Fλ ranges from −3.5% to −0.7%,
with per image standard deviations ranging from 6% to 10%. As discussed above, we also
found a slight negative error when we compared the DAOPHOT photometry to aperture
5
Figure 1: The process by which we remove the background and mask the nucleus in MRI images. (a) A
fully calibrated image, mv5004046. (b) The median-filter subtracted image. (c) Image of the morphological
gradient of image b. (d) The final mask, defined where image c > 0.1 W m−2 µm−1 sr−1. The mask has
been dilated by 30 pixels to fill gaps and cover some additional coma. The image scale and orientation are
given in panel d. The projected vectors are comet-Sun ((cid:12)), comet-Earth (⊕), and spacecraft velocity (vDI).
6
Figure 2: Example point source detection completeness functions. All images were taken within ∆ =
694 − 704 km, except mv5004009 (501 ms) taken at ∆ = 1858 km. The mv5004046 sub-frame is chosen to
match the same area as covered in hv5004031. The HRIVIS test particles are selected from one of 9 specific
fluxes (marked by squares and circles), whereas the MRI test particles are selected at random from a range
of fluxes. The lowest of the HRIVIS bins is 10,000 e−.
7
photometry of isolated bright sources. To account for this effect, we will increase the fluxes
of all particles by 2%. In addition, we adopt 8% as our particle flux uncertainty.
The large flux errors (∆F/F > 10%) are likely misidentified artificial particles, artificial
particles that have merged with brighter particles in the PSF fitting process, and PSF
over/under fitting. Any particles with significantly large flux errors must be identified and
removed from our photometry lists so that they do not affect the final results. We clean the
DAOPHOT output by rejecting: (1) bad PSF fits as identified by DAOPHOT (i.e., reduced
χ2 > 2.0); and (2) any particles with a total residual in a 3 × 3 pixel box greater than 3σ
from the background. For the 41 ms images 62 − 78% and 0 − 2% of the detected sources
were rejected by the two criteria, respectively. For the 121 ms images the rejection rates were
5−32% and 2−5%; for 501 ms, the rates were 0−1%, and 3−20%. Occasionally DAOPHOT
fits portions of the particle streaks, and these criteria help remove such spurious fits from
the photometry lists. Examples of point source cleaned images are shown in Figs. 4 and 5.
After point source cleaning, sources not fit by DAOPHOT are evident. These particles can
be accounted for in a statistical sense with the image's completeness function, e.g., when we
are displaying the flux distribution of the particles, we divide the measured flux distribution
by the completeness curve to show the true flux distribution.
In Fig 5 we specifically show a region of high point source density, and the large number of
bad-PSF fits resulting from point source crowding near the nucleus (> 0.025 pixel−1). These
sources are primarily rejected by the χ2 criterion. Regions with a rejection rate > 20% in a
25 × 25 pixel area will be masked from our analyses.
3.2. HRIVIS images
The analysis of the HRIVIS images required an approach independent from the MRI
analysis. Figure 6a shows an HRIVIS image of the large particles and Fig. 6b is the restored
(i.e., deconvolved) image. The broad defocused PSF and the longer trails resulting from
parallax motion make PSF fitting even more difficult than in the MRI images. Our approach
is to instead detect and measure a particle's brightness in restored images. In summary, we:
(1) iteratively remove the background from each image; (2) detect point sources and streaks
in the restored images; (3) measure particle fluxes; and, (4) measure the completeness of our
particle detection scheme. Below we detail our methods.
We use an iterative technique to remove the background before final object detection
and photometry. The initial background is set to the deconvolved image filtered with a
51× 51 pixel median filter. We then fit a smooth surface (using cubic splines) to the filtered
image to obtain our background image. The objects within the image will bias the median
filter and the resulting image overestimates the background. To minimize this bias, we repeat
the process after ignoring all pixels more than 400 e− above the background (object pixels)
during median filtering (1 e− s−1 = 4.32 × 10−6 W m−2 µm−1 sr; Klaasen et al. in prep.).
We repeat this process three times to obtain and remove our final background (Fig. 6c).
We detect the objects in the image by creating a mask of all pixels more than 1,000 e−
above the background level (Fig. 6d). We selected the 1,000 e− threshold by visual inspection
of the restored image to avoid detection of restoration artifacts that result from "ringing"
around the brighter objects (Lindler et al. this issue). We define contiguous masked pixels
as part of the same object. Visual inspection of this mask shows many instances where a
trailed particle near the detection threshold is split into multiple objects because of noise
8
Figure 3: Input flux, Fλ, plotted versus DAOPHOT error, ∆Fλ/Fλ, for all artificial particles inserted in
image mv5004046. Most fluxes are measured to within 8% (1σ) of their true value, with a slight skew toward
negative errors.
9
Figure 4: (a) A detail of image mv5004046, after background coma subtraction. (b) The same image after
being cleaned of point sources detected and fit by DAOPHOT. In both images, particles with large residuals
or low signal-to-noise flux ratios have been circled. Similar to Figure 1, the image scale and orientation are
shown. The square in the lower right image shows the field of view of panels a and b with respect to the
nucleus.
10
Figure 5: The same as Fig. 4, but for a region of high point source density in image mv5004029.
11
Figure 6: (a) Original HRIVIS image hv5004030. (b) Image deconvolved with the R-L method. (c) Image b
after background subtraction. (d) Mask of pixels greater than 1000 e− above the background. (e) Mask of
detected sources. (f) Mask of the sources used to compute the flux distribution slope (here, length 5 pixels
or smaller and total flux greater than 20,000 e−).
12
and potential variations in particle brightness during the time of the exposure. To avoid this
problem, we examine each detected object (starting with the brightest) and adding all pixels
within a factor of 3 of the brightest pixel in the object. We define the resulting contiguous
dark regions as our detected objects (Fig. 6e).
We perform photometry on the detected particles by starting with the particle mask, and
adding the flux from their nearest-neighbors (i.e., we grow the mask by 1 pixel, and sum
the pixels together). Non-linearities are a concern when restoring an image with non-linear
deconvolution algorithms (Lindler et al. 1994). Fainter point sources will be broader than
brighter point sources in a restored image. We can minimize any non-linearities by increasing
the size of the region used for photometry. However, this results in more contamination of our
results by nearby sources. To test for non-linearity we increased the size of the photometry
regions by adding additional neighbors (i.e., including next-nearest neighbors, etc.). These
larger regions showed that non-linearity artifacts are absent from particles brighter than
10,000 e− (Fλ = 1.4 × 10−12 and 4.6 × 10−13 W m−2 µm−1 for exposure times of 126 and
376 ms). On average, the effective aperture for point sources is approximately a 3 pixel
radius circle.
As done with the MRI analysis, we measure the completeness by inserting artificial
objects in the images of varying brightness and parallax length. These artificial objects are
added to the raw images before deconvolution. The completeness test also gives additional
evidence that our measured fluxes are linear with object brightness in the range of brightness
being considered. For the longer trails, the completeness decreased rapidly with flux. We
therefore decided to ignore particles with a length more than 5 pixels in the restored image.
In effect, this criterion limits the line-of-sight distance of the particles from the nucleus to
< 1 − 2 km at closest approach (∆ = 694 km), and < 2 − 3 km at ∆ = 900 km. If the flux
distribution is independent with distance from the nucleus, then the relative distribution
of number versus flux will remain unchanged. Limiting the length of the objects also has
added benefits. Long streaks are much more likely to be a combination of multiple particles
and their photometry is also more strongly affected by errors in the background estimation
(they cover more pixels). Examples of our HRIVIS completeness test results are presented
in Figure 2. There is little difference between the 126 ms and 376 ms completeness curves in
the example. Therefore our completeness is not as strongly affected by background noise as
it is by other factors, e.g., point source crowding and deconvolution artifacts. In addition,
our HRIVIS photometry is generally limited by the linearity criterion of F > 10, 000 e−.
4. Particle fluxes
4.1. Flux distribution
We measure the fluxes of all particles and compute their flux distributions, dn/dF . The
particle fluxes approximately follow a power-law distribution. Therefore, we fit each flux
distribution with a function of the form dn/dF ∝ F α using the method by Clauset et al.
(2009), which is based on Kolmogorov-Smirnov tests and maximum likelihood fitting. We
restricted the flux range to particles brighter than our 80% completeness estimate. This
restriction increases the statistical errors in the fits but reduces possible bias errors resulting
from photometry contamination from neighboring particles and residual background. More-
over, the fitting method does not include a correction for completeness, so restricting the
13
fluxes helps mitigate the effect of incomplete counting; the maximum completeness correc-
tion over our fit range is 20%, which is insignificant in comparison to the strong power-law
slopes (see Fig. S6 of A'Hearn et al. 2011). Our results are listed in Table 2 and summarized
in Table 3. In Figs. 7 and 8, we plot the final flux distributions and their best-fit trends.
The fluxes are plotted as a function of Fλ,700 ≡ Fλ∆2/7002 so that they may be directly
compared to each other (variation of the phase angle is small, ranging from 79 to 92◦, and is
therefore not considered). The error-weighted mean slopes for the MRI and HRIVIS frames
are −3.74 ± 0.02 and −2.85 ± 0.03, respectively.
The MRI and HRIVIS slopes are significantly different from each other. To verify this
difference, we also computed the flux distributions from MRI sub-frames of images chosen
close in time to the HRIVIS images. The relative position of the HRIVIS and MRI fields
are constant with time, therefore measuring the flux distribution in the same field of view is
straightforward. The comparison, however, is not exact because the frame-to-frame parallax
of individual particles can be significant (Hermalyn et al. this issue). After bad-PSF rejection
and completeness tests only ≈ 130 particles could be used to determine the flux distributions
in the MRI sub-frames. This is a factor of 3 -- 10 fewer than the number of particles in the
HRIVIS images. The difference primarily relies on the larger aperture of the HRI primary,
which allows for a fainter point source detection limit due to the finer resolution and greater
light gathering area. Our best-fit parameters for the MRI sub-frames are listed in Table 2 and
summarized in Table 3. The uncertainties in the power-law slopes are quite large, ranging
from 0.18 to 0.31, owing to the paucity of particles available in the MRI sub-frame images.
The weighted mean slope is −3.59 ± 0.07, which is shallower than the mean MRI full-frame
slope, but they agree at the 2σ level.
Taken together, the mean MRI full-frame, MRI sub-frame, and HRIVIS best-fit power-
law slopes show evidence that the flux distribution steepens with increasing flux. Inspection
of the MRI full-frame distributions in Fig. 7 and the Kolmogorov-Smirnov probability (PKS)
values reveals that a power-law size distribution is not a good fit to most of the flux dis-
tributions. This poor fit is especially true for the 121 ms frames at ∆ < 900 km where
there is a significant curvature in the log-flux distributions of these images. The curvature
is non-existent or not prevalent in the shorter exposure MRI data, the distant MRI data
(∆ > 1000 km), and the HRIVIS data.
It is important to recognize two points. First, the HRIVIS distributions are poorly
populated at Fλ,700 > 10−11 W m−2 µm−1, whereas the 121 ms MRI distributions are well
defined up to Fλ,700 ≈ 3 × 10−11 W m−2 µm−1. This difference is simply a matter of the
larger solid angle observed by the MRI camera as compared to the HRIVIS camera; a larger
solid angle allows for more of the brightest particles to be detected. Second, the HRIVIS
distributions are valid to lower fluxes than the MRI distributions (column Fλ,min in Table 3),
because the larger primary of the HRI allows for fainter particles to be accurately measured.
Taken together, these differences demonstrate that the two instruments are measuring two
different particle flux regimes. In Fig. 9, we plot our best-fit power-law slopes versus the
minimum flux used in each fit. A trend is evident; steeper slopes are correlated with larger
Fλ,min, but due to the different techniques, fit uncertainties, and fields of view involved,
the correlation is not conclusive. We remind the reader that the correlation is not due to
incomplete photometry because we restricted our data sets to fluxes where the completeness
is better than 80%. Given this restriction, none of our completeness corrections are strong
14
Figure 7: Flux distributions (dn/d log F ) and their best-fit power-law functions for all full-frame MRI data
sets listed in Table 2. The flux distributions have been corrected for completeness, and their error bars are
based on Poisson statistics. The flux distributions are grouped by their integration times, and sorted by
time with the pre-closest approach images at the bottom. Each i-th flux distribution (i = 0, 1, . . .) has been
offset along the y-axis by 10i, and labeled with ∆ for clarity. Additionally, the fluxes have been scaled by
(∆/700)2 so that they are directly comparable to each other. The gray × symbols mark particles not used
in the fit.
15
Figure 8: Same as Fig. 7, but for all HRIVIS images listed in Table 2.
16
Figure 9: Best-fit power-law slope versus minimum flux used in the fit for all images listed in Table 2.
enough to have a significant effect on the best-fit slopes.
In the absence of any additional information on the flux distribution we will proceed
assuming that the differences between the MRI and HRIVIS flux distributions reflect the true
flux distribution of the entire particle population. As alternatives to a power-law distribution,
we also considered a power-law with an exponential cut off, and a double power-law. The
functional forms are
dn
dF
dn
dF
β2(cid:18)
∝ F β1 exp (−F/γ1) ,
∝ F
γ2
(cid:19)β3−β2
1 +
F
γ2
,
(1)
(2)
where βi are power-law slopes, and γ1 is the flux at which the exponential decay reaches
a factor of exp (−1) = 0.37 , and γ2 is the turn-over flux from the low- to the high-flux
slopes (both γ parameters are specified for ∆ = 700 km). The double power-law fits did not
converge on a single solution. The power-law with exponential cut-off function, however, is
a better fit to the closest-approach data (∆ < 1000 km), with a mean slope and cut-off equal
17
to −3.18 ± 0.07 and (5.7 ± 0.3) × 10−12 W m−2 µm−1. Unfortunately, these best-fit values
are poor fits at larger distances where the flux distribution better agrees with a power-law
(Fig. 10). We suggest that the 121 ms distributions have a systematic effect that causes
under-counting of the largest flux bins. Again, we note that our completeness corrections
are not strong enough to significantly affect our fits, and that the distributions as shown in
Figs. 7 and 10 have been corrected for completeness, yet the curvature remains in the closest
approach data.
To test the hypothesis that the power-law slope is affected by the source density, we
generated completely synthetic data sets with DAOPHOT, based on our MRI PSF. We
created 8 images each with the same power-law flux distribution (dn/dF ∝ F −3.5) and
maximum point source flux (Fλ,max = 1 × 10−10 W m−2 µm−1), but Fλ,min varied from
5×10−12 down to 1.5×10−13 W m−2 µm−1. We inserted the same number of point sources per
logarithmic flux bin, but because each image's flux distribution spanned a different range, the
total point source density varied from 4× 10−4 pixel−1 up to 2.6 pixel−1. We then measured
each image's flux distribution with DAOPHOT in a manner similar to our MRI images.
Once the input column density reached 0.08 pixel−1, our best-fit slopes steepened from the
input −3.5 down to a minimum of −4.7 at the highest densities. Thus, the rollover in the
flux distribution seen in the 121 ms MRI images at ∆ < 1200 km could be a manifestation
of this effect.
The slopes derived from the distant MRI and the 41 ms MRI images are consistent:
α = −3.80 ± 0.02 and −3.83 ± 0.05, respectively. They agree despite the wide range in the
number of particles fit (100 -- 1000) and spacecraft-comet distances (900 -- 5000 km, although
the best constraints are within 3000 km). Therefore, we consider these results to be robust
and representative of the true flux distribution for Fλ,700 (cid:38) 1 × 10−11 W m−2 µm−1. The
shallower HRIVIS slopes appear valid for 1 × 10−12 (cid:46) Fλ (cid:46) 1 × 10−11. We see no reason to
trust one data set over the other, or to assume that a single power-law would be valid for all
fluxes measured. Therefore, we will use a broken power-law for the remainder of the paper:
(cid:40)
∝ F α
dn
dF
α = −2.85
α = −3.80
for Fλ,700 < 7.2 × 10−12 W m−2 µm−1
for Fλ,700 ≥ 7.2 × 10−12 W m−2 µm−1
,
(3)
where Fλ,700 = 7.2× 10−12 is the break in the power-law, −2.85± 0.02 is the average HRIVIS
slope, and −3.80 ± 0.02 is the average MRI slope, measured from images at ∆ > 900 km
to avoid the apparent crowding effects at closest approach. The break was derived by least-
squares fitting the combined MRI and HRIVIS data sets. The broken power-law and the
combined HRIVIS and MRI flux distributions are presented in Fig. 11. This model is a good
match to the ensemble (HRIVIS + MRI) data set.
4.2. Total particle flux
Due to the great numbers of large particles, we can only accurately measure the total
scattered flux from the brightest particles in each image. Our best estimate of the frac-
tion of the coma flux attributable to large particles is 0.11% for fluxes ranging 0.9 − 4.7 ×
10−11 W cm−2 µm−1 (row labeled "MRI (distant)" in Table 3). This estimate was derived
from images with relatively little point source crowding (∆ > 2000 km), and a moderate
18
Figure 10: Best-fit power-law with exponential cut off, derived from 121 ms MRI images taken within
∆ = 1000 km, compared to selected flux distributions, labeled with "∆ (exposure time)" in km and ms. The
histograms have been offset by a factor of 10i for clarity.
19
Figure 11: Combined HRIVIS and MRI flux distributions (Fλ ≥ Fλ,min), scaled to N (Fλ = 7.2× 10−12) = 1.
For the MRI data, only full-frame distributions from ∆ > 900 km are shown. The solid line is our best-fit
broken power-law (Eq. 3).
20
number of particle measurements (N > 300 particles). In order to build a complete cen-
sus of the particles, we need to extrapolate the results from that limited range down to
the faint end of the distribution. We find that the brightest particles observed throughout
the closest approach images are consistently near Fλ,700 = 4.5 × 10−11 W cm−2 µm−1. On
the other end, the faintest particles we can manually find and measure have fluxes of order
Fλ,700 ∼ 10−13 W cm−2 µm−1 (Fig. 12). If the very steep flux distribution we derived in §4.1
extends from Fλ,700 = 1 × 10−12 down to 1 × 10−13 W cm−2 µm−1, there should be several
million particles per image. However, at such great pixel densities (1 particle per 17 pixel
area) it should be very difficult to find faint isolated particles (the core of the HRIVIS native
PSF has an area of 113 pixels). Furthermore, if we let the flux distribution continue down
to 1× 10−14 W cm−2 µm−1, the large particles account for 100% of the coma flux, leaving no
room for any fainter particles (millimeter sized and smaller). Therefore, the flux distribution
must change or be truncated at fluxes fainter than Fλ,700 = 1 × 10−12 W cm−2 µm−1. The
uncertainty in the lower flux limit is a major source of error in all estimates of the total
number, flux, cross section, and similar derived quantities. We list the results from our ex-
trapolations in Table 4. We estimate that the large particles account for 2 -- 14% of the total
coma flux near the nucleus.
If we take the the first image in which particles are easily seen, mv5002051 at ∆ = 5148 km
(Fig. 13), we can estimate the total number of particles within a 20.6 km radius (the largest
circular aperture centered on the nucleus that fits within the image). We find a total coma
flux of (2.15 ± 0.22) × 10−7 W m−2 µm−1 (includes the bright jets, but not the nucleus) and
assume that 2 -- 14% of that flux is attributable to large particles. The results are presented
in Table 4 and will be used below when we estimate the cross section, mass, and water
production rate of the large particles.
Up to now, we have not addressed streaked particles in our MRI images. If a significant
fraction of the flux in our images at ∆ ≈ 2400 km is contained within streaked particles
then our estimates on the total flux and number of large particles will be systematically low.
However, we find that few particles are streaked and the correction to include any possible
streaks is negligible. To demonstrate, we developed a Monte Carlo simulation that uses the
spacecraft position and velocity to estimate the volume of space that is smeared. Within a
20.6 km projected radius, 2% the field-of-view is smeared over ≥ 1.5 MRI pixels, and only
If the particle density follows a r−2
particles outside of 120 km are smeared this much.
profile, the fraction of smeared particles reduces to effectively zero.
5. Particle size and composition
In the absence of any compositional information on these large particles, we assume two
cases to demonstrate their likely range of sizes. First, we will consider that the particles are
refractory, and photometrically behave like comet nuclei (the "dusty" case). Then, we will
consider that the particles are icy, and behave like the icy satellite Europa (the "icy" case).
We stress that these two cases are examples only. They may not reflect the true nature
of the particles, but they do yield useful limits on the particle sizes. We will show that
the particles are much larger than the 0.1 -- 100 µm sizes typically considered in comet dust.
We purposefully avoid interpreting the large particles with phase functions that have been
derived for comet comae. Light scattered by comet dust comae is expected to be dominated
21
Figure 12: Two sub-frames of HRIVIS image hv5004031 (∆ = 694 km) showing the crowded field of particles:
a) an MRI context image (mv5004046) with the approximate HRIVIS field of view outlined with a box (image
is log scaled from 0.001 to 1.0 W m−2 µm−1 sr−1); b) HRIVIS image, sub-frames c/d and e/f are outlined and
labeled; c) HRIVIS sub-frame; d) the same as c, but deconvolved to enhance the spatial resolution; e) another
HRIVIS sub-frame; f) panel e, deconvolved. Two particles with fluxes near 1− 2× 10−13 W m−2 µm−1 have
been circled. The HRIVIS field of view is 1.4 km, and the sub-frames are 0.42 km.
22
Figure 13: Image mv5002051, background coma removed. Point sources identified by DAOPHOT with fluxes
greater larger than the image's 80% completeness limit are circled to show the extent of the large particle
coma (particles below this limit can still be seen by eye). An image of the nucleus has been inserted into
the central masked region. The box marks the location of the inset image. Stars, most of which have been
automatically masked, are streaked over 13 pixels in this image.
23
by dust grains in the sub-micrometer to micrometer size range (Kolokolova et al. 2004),
which scatter optical light differently than centimeter-sized particles.
First we take the case in which the particles photometrically behave like comet nuclei,
i.e., they have a very low albedo, and have a phase angle behavior like a macroscopic object
and not like small dust grains. We refer to this case as the "dusty case," but, just like
comet nuclei, these model particles may be internally icy. We adopt the geometric albedo
(Ap = 0.049) and phase function (−2.5 log Φ = 0.046θ mag for θ in units of degrees) of
Hartley 2's nucleus (Li et al. this issue). The geometric albedo is defined as the ratio of
the energy scattered from the object toward a phase angle of 0◦ to that scattered from a
white Lambertian disk with the same cross section (cf. Hanner et al. 1981 for this and other
relevant albedo definitions).
In Section 4.1, we found that individual particle fluxes range from Fλ,700= 1 × 10−13 to
4.5×10−11 W m−2 µm−1, where Fλ,700 is the particle flux normalized to a distance of 700 km.
To convert between flux and cross section, we use the formula
Fλ = ApΦ(θ)σ
Sλ,(cid:12)
r2
h
1
∆2
(4)
where Fλ is the particle flux in units of W m−2 µm−1, σ is the cross-sectional area of the
particle (cm2), Sλ,(cid:12) is the solar flux density at 1 AU (W m−2 µm−1), rh is the heliocentric
distance of the particle (1.064 AU), and ∆ is the spacecraft-particle distance (cm). For the
HRIVIS and MRI CLEAR1 filters, we use a solar flux density of 1471 and 1435 W m−2 µm−1,
respectively (Klaasen et al. in prep.). To convert from cross section to effective radius,
we assume a spherical geometry: σ = πa2. Altogether, assuming a comet nucleus-like
photometric behavior, the effective radii of the particles range from 10 to 221 cm. The largest
of these particles (4 m diameter) are just over the resolution of the HRIVIS reconstructed
frames (about 3 m at a distance of 700 m). We consider these sizes to be an upper limit to
the true particle sizes. For a lower-limit estimate, we consider the case where the particle
scattering function is similar to the icy satellite Europa: Ap = 0.67 and Φ = 1.0 − 0.01θ +
2.2 × 10−5θ2 (Buratti and Veverka 1983, Grundy et al. 2007). At a phase angle of 80◦,
Europa's phase function is 0.34. For this parameter set the effective radii are 0.8 to 17 cm.
Following Eq. 4, where flux is proportional to cross-sectional area, we can compute the
total observed particle cross section and add it to Table 4. We have assumed our icy particle
case, using the photometric parameters of Europa. To instead use our dusty case, multiply
the cross sections in the table by 158.
Our flux distributions imply a very steep size distribution. Assuming a spherical geome-
try, the differential size distribution is
dn
da
= 2N0F α+1
1
a2α+1,
(5)
where F1 is the flux from a 1 cm radius particle in units of W m−2 µm−1, and particle radius
a is in units of cm. The constant N0 is the solution to the equation:
(cid:90) Fλ,max
Fλ,min
24
N = N0
dn
dF
dF,
(6)
using the appropriate values from Table 4. Our low-flux power-law slope (α = −2.85)
corresponds to a size distribution proportional to a−4.7. This slope is steeper than the many
size distribution estimates of small through large dust grains (a ∼ 1 µm to 1 mm with slopes
near −4 to −3) based on grain thermal emission (Lisse et al. 1998, Harker et al. 2002, 2011),
grain dynamics (Fulle 2004, Reach et al. 2007, Kelley et al. 2008, Vaubaillon and Reach 2010),
and dust flux monitors on spacecraft (McDonnell et al. 1987, Green et al. 2004). Specifically
for Hartley 2, Bauer et al. (2011) estimate the comet's overall size distribution to follow a
power-law slope of −4.0± 0.3, derived by comparing R-band and WISE 12 and 22 µm fluxes.
Epifani et al. (2001) fit an ISOCAM 15 µm image of the comet taken about 10 days after
perihelion with a dust dynamical model. They report a time-averaged power-law slope of
−3.2 ± 0.1, but inspection of their Fig. 10 suggests −3.8 is more appropriate (their best-fit
slopes never fall above −3.5). These slopes are shallower than our value of −4.7, but there
is no requirement that they be the same.
To better understand the size, and thereby the composition, of the large particles, we
investigate the largest particle that may be lifted from the nucleus, acrit. This parameter is
estimated by comparing the force of gravity to the gas drag force at the surface of the comet.
Meech and Svoren (2004) integrated the equation of motion for spherical particles ejected
from a spherical nucleus and found
acrit =
9µmHQvth
64π2ρpρN R3
N G
,
(7)
3µmHQvth
16πρpaN R2
N
.
where µ is the atomic weight of the gas (amu), mH is the mass of hydrogen (g), Q is the gas
production rate (s−1), vth is the mean thermal expansion speed of the gas (cm s−1), ρp is the
density of the particle (g cm−3), ρN is the density of the nucleus (g cm−3), RN is the radius of
the nucleus (cm), and G is the gravitational constant (cm3 g−1 s−2). With the shape model
of Comet Hartley 2, Thomas et al. (this issue) estimate the surface gravity of the nucleus to
be aN = 0.0019− 0.0044 cm s−2, which includes the rotation state of the nucleus. Therefore,
we re-write the equation from Meech and Svoren (2004) to use the gravitational acceleration
at the nucleus
acrit =
(8)
The bulk material density for dust is ∼ 3 g cm−3 and for ice is 1.0 g cm−3. For the dusty
case, we will consider porous aggregates of dust with a total density of 0.3 g cm−3 (i.e., 90%
vacuum). For the icy case, we will at first assume 1.0 g cm−3, but later consider porous
aggregates with ρp = 0.1 g cm−3.
The total water production rate of Comet Hartley 2 near closest approach has been
estimated via several methods to be Q(H2O) ≈ 0.7 − 1.2 × 1028 s−1 (A'Hearn et al. 2011,
Combi et al. 2011, Dello Russo et al. 2011, Meech et al. 2011, Mumma et al. 2011, Knight
and Schleicher this issue). Yet, the coma contains a significant amount of water ice (A'Hearn
et al. 2011, Protopapa et al. 2011), which could be supplying a large fraction of the water
vapor around the comet. Moreover, the water production rate is not uniformly distributed
over the surface (A'Hearn et al. 2011). We can account for these observations by multiplying
the water production rate by the ratio fsurf ace/factive, where fsurf ace is the fraction of the
water vapor produced at the surface of the nucleus, and factive is the areal fraction of surface
25
that is active. For illustrative purposes, we will assume that 1% of the water vapor is
produced from 10% of the surface. The remaining 99% of the water vapor sublimates from
the icy grain halo.
For a nucleus water production rate of Q(H2O)fsurf ace/factive = 1027 s−1, vth = 0.5 km s−1,
Hartley 2's mean radius (0.58 km), mean surface gravity, and assuming a particle density of
pure ice (ρp = 1 g cm−3), we find acrit = 8 cm. If we instead take CO2 as the driving gas,
with a production rate of 1027 s−1 (A'Hearn et al. 2011), fsurf ace = 100%, and factive = 10%,
then acrit becomes 200 cm for solid ice spheres. Based on this exercise, the icy model size
estimates, a ≤ 17 cm, are reasonable.
If instead of icy particles, we assume nucleus-like particles with a density of 0.3 g cm−3,
our acrit estimates increase to 28 cm (H2O) and 670 cm (CO2). Compared to our size
estimates of a ≤ 210 cm, dark, dusty particles are plausible if CO2 is the driving gas; water
can be made consistent if fsurf ace/factive is increased to 1.0.
Aggregate particles are more easily lifted by gas drag due to their larger surface area
per mass. Nakamura and Hidaka (1998) found that the drag force for an aggregate is
approximately the same as the drag force on an area-equivalent sphere (with an error of less
than 40% in the large aggregate limit). Since our particle radii are based on the observed
flux, which is proportional to the cross-sectional area, our radii are already defined by area-
equivalent spheres. Therefore, by revising our particle density we can use Eq. 8 to estimate
acrit for aggregates.
We assembled model particles using a ballistic particle-cluster aggregate (BPCA) method
(Meakin 1984). As the size of the aggregate grows beyond a few thousand monomers the
density asymptotically approaches 10% of the bulk material density. Thus, for a refractory
material with a bulk density near 3 g cm−3, a centimeter-sized BPCA particle would have a
density of 0.3 g cm−3. The acrit estimates will be the same as in our nucleus-like case above.
Treating the large particles as aggregates rather than solid spheres better agrees with the
HRIIR spectra of the coma. A'Hearn et al. (2011) and Protopapa et al. (2011) studied the
water ice absorption features and found that they are most consistent with icy aggregates
with monomer radii ≈ 1 µm. An icy BPCA particle would have a density of 0.1 g cm−3. The
lowered density for icy aggregates increases acrit by a factor of 10, giving us a healthy margin
for launching large icy particles off the surface of the nucleus, even where water sublimation
is driving the activity.
In summary, the dusty particle case produces very large particle estimates (up to 2 m
in radius) that are just at the resolution limit of the HRIVIS instrument. Gas expansion
from CO2 is sufficient to lift these large particles from the surface of the comet if they have
a comet-like density of ≈ 0.3 g cm−3. The water-ice case produces particle estimates up to
≈ 20 cm in radius, which are easily lifted from the nucleus by water or CO2 expansion.
6. Spatial distribution and origin
Mapping the spatial distribution of the particles gives clues to their origins and dynamics.
In Figs. 14 -- 17, we plot the column density of particles and total coma surface brightness
contours for all 501 and 121 ms MRI images listed in Table 2. The column density images
were derived from our final photometry lists for Fλ > Fλ,min, binned onto a 38 × 38 pixel
grid. Inspection of the figures reveals that the coma and the particles have different spatial
26
distributions. The particles are biased to the anti-sunward direction on scales > 2 − 4 km,
whereas the coma distribution is dominated by the jets. This asymmetry is especially ap-
parent in the strong sunward jets, where the particle density is lower than in other regions
of similar surface brightness. It is not an observational bias; we have masked those regions
close to the nucleus where the column density and jet morphology interfere with the PSF
fitting process. Instead, the low particle density in the sunward jets can be accounted for
by particle dynamics. We consider three dynamical processes that could be affecting the
particle distribution: (1) the rotation of the nucleus; (2) solar radiation pressure; and (3)
a rocket effect from sublimating ice. Hydrodynamic flows from the strong gas production
rates and asymmetry in outgassing may also play a role in the particle dynamics, but this
analysis is outside the scope of this paper.
6.1. Radial expansion and nucleus rotation state
Rotation of the nucleus has a direct impact on the spatial distribution of particles. To
estimate this effect, we must first recognize that the particle outflow speeds are low.
In
order for the particles to be seen as point sources in HRIVIS images at closest approach,
their speeds must be lower than ≈ 3 m/0.376 s = 8 m s−1. A lower constraint is computed by
Hermalyn et al. (this issue) based on the 3D positions of the particles. They find 0.5 -- 2 m s−1
to be more typical (note that these speeds are not necessarily radial). At such low speeds,
the large particles take (cid:38) 103 s to reach 2 km from the nucleus. In contrast, small dust
grains move much more quickly with outflow speeds expected to be of order 100 m s−1. The
fine dust reaches 2 km in as little as 20 s.
The positions of the major jets are governed by the rotation of the long axis about the
angular momentum vector, with a period of 18.4 h near closest approach; the long axis is
inclined to the angular momentum vector by 81◦ (Belton et al. this issue). Taking 1 m s−1
as the particle outflow speed from the surface, the nucleus will have rotated 10◦ by the time
the particles have traveled 2 km. So, the rotation state has a minor consequence on the
distribution of particles at 2 km, and their spatial distribution should be closely related to
their source regions. This observation and the assumption of radial motion suggests that
the strong sunward jets are not the primary source of the large particles, but instead they
are ejected from along the long-axis of the nucleus. The jets pointed towards the bottom
of the images in Fig. 15 would be the next likely source region for particles. There is also
a large population of particles towards the top of Figs. 14 and 15, but they do not have an
apparent source region (i.e., this side of the nucleus does not appear to be as active as the
other regions). However, we note that not all potential sources are apparent in the MRI
images. For example, the apparent water jet seen in Fig. 5 of A'Hearn et al. (2011), which
points to the top right of Figs. 16 and 17, has no clear optical counterpart but may also
contribute to the large particle production (n.b., this water jet does not appear to contain
ice, and therefore is unlikely to be a source of large icy particles).
Taking a lower ejection speed of 1 cm s−1, the nucleus can rotate three times before
particles travel 2 km. Thus on this length scale, and in the absence of any other perturbing
forces, the particles would have a spatial distribution correlated with the activity profile of
the nucleus. Because the CO2 and H2O gas production rates and the optical light curve peak
when the small end is pointed towards the Sun (A'Hearn et al. 2011), in the absence of other
forces the large particle density should peak in the solar direction, which is not observed.
27
Figure 14: Contours of the total coma surface brightness superimposed over particle column density for 501
and 121 ms MRI images mv5002051 through mv5004025. The contours are spaced at factor of 2 intervals,
and the dashed line is 0.008 W m−2 µm−1 sr−1. The coma image was smoothed with a Gaussian kernel
function (7 pixel FWHM) before the contours were created. Only particles with Fλ > Fλ,min are considered,
and the particle bins are 38× 38 pixels in size. The region closest to the nucleus is masked from the analysis,
and has been replaced with each epoch's image of the nucleus and inner-most coma.
In all images, the
projected velocity of the spacecraft is toward the bottom, and the sunward direction is approximately to the
right.
28
Figure 15: Same as Fig. 14, but for images mv5004029 through mv5004053.
29
Figure 16: Same as Fig. 14, but for images mv5004056 through mv5006020.
30
Figure 17: Same as Fig. 14, but for images mv5006024 through mv5006046.
31
6.2. Solar radiation pressure
In principle, solar radiation pressure could redistribute large particles into the anti-
sunward direction. The acceleration from radiation, arad, in units of cm s−2 is (Burns et al.
1979)
QprS(cid:12)σ
cmr2
h
,
arad =
(9)
where Qpr is the radiation pressure efficiency factor, S(cid:12) is the integrated flux density (1.361×
106 erg cm−2 s−1 at 1 AU), σ is the geometrical cross section of the particle (cm2), c is the
speed of light (cm s−1), m is the mass of the particle in question, and rh is the heliocentric
distance (AU). We have already dropped all velocity dependent terms from arad (cf. Burns
et al. 1979). It is common to express radiation pressure with the parameter β defined as the
ratio of the force of solar radiation pressure to the gravitational force from the Sun,
β ≡ Frad
Fgrav
=
5.7 × 10−5Qpr
ρpa
.
(10)
(cid:82) π
The radiation pressure efficiency is 1 for perfectly absorbing, isotropically emitting spheres.
For our particles, we again take the icy particle case and derive Qpr from the phase func-
tion and geometric albedo (van de Hulst 1957): Qpr ≈ 1 − ABcos α, where AB is the Bond
albedo, and cos α describes the anisotropy of the scattered light, where α = 180 − θ is
the scattering angle. The relationship between the Bond albedo and geometric albedo is
AB = Ap
0 Φ(α) sin αdα (Hanner et al. 1981). Altogether, we compute Qpr = 1.58. For the
mass, we again assume two cases for particles with a = 10 cm: (1) solid spheres with the
density of ice, 1 g cm−3; and (2) compact aggregates of ice (1 µm radius monomers) with an
overall particle density of 0.1 g cm−3. For the latter case, we note that detailed calculations
will be required to understand how albedo and the anisotropy of scattering are affected by
the complex aggregate shape, and we reserve this investigation for future work. We compute
accelerations of 4.8 × 10−6 and 0.5 × 10−6 cm s−2 for 10 cm solid and BPCA aggregate ice
particles, equivalent to β = 9 × 10−6, and 1 × 10−6. For comparison, comet dust trails are
comprised of grains with β (cid:46) 10−3 (Sykes and Walker 1992, Reach et al. 2007).
In Figs. 14 -- 17, particles are found out to the image edges in the sunward direction.
However, the sunward/anti-sunward asymmetry is clear on 2 -- 4 km length scales. In the rest
frame of the comet, the relationship between turnaround distance (d), ejection speed (vej),
and acceleration from radiation pressure is d = v2
ej/2/arad. Solving for ejection speed vej
yields
3dQprS(cid:12)
2cr2
hρpa
= 6.0 × 10−5 dQpr
ρpa
v2
ej =
,
(11)
for vej in units of cm s−1, d in cm, ρp in g cm−3, and a in cm. For our icy particle case
vej = 4.3−6.1(ρa)−1/2 cm s−1 (4 -- 6 cm s−1 for a 10 cm aggregate, and 1 -- 2 cm s−1 for solid ice).
If particles are truly ejected at these speeds, solar radiation pressure would take (cid:38) 106ρa s to
accelerate the particles to the (cid:38) 50 cm s−1 speeds measured by Hermalyn et al. (this issue).
This result also implies that the dominant velocity component for particles more than a few
kilometers from the nucleus would be distinctly in the anti-solar direction, yet only a weak
asymmetry in the velocity is observed (Hermalyn et al. this issue). Therefore, radiation
pressure does not govern the dynamics for our icy particle cases. The same conclusion is
32
reached for the dusty particle case with a nucleus-like density (Qpr = 1.00, ρp = 0.3 g cm−3):
vej = 6.3 − 8.9a−1/2 cm s−1 or 0.6 -- 0.9 cm s−1 for a 100 cm particle.
If we instead assume an ejection speed of 1 m s−1 for a 10 cm particle, their densities
must be ρp = 2 − 4 × 10−4 g cm−3 in order to be turned around by d = 2 − 4 km. Thus, if
the particles are very fluffy aggregates, they may be ejected at larger speeds, and radiation
pressure will be able to redistribute them into the anti-solar direction, but they could still
have instantaneous velocities distributed about the anti-solar vector.
6.3. Rocket effect
The rocket effect is the acceleration of the particles due to the sublimation of water ice.
For a spherical geometry with radial outflow,
arocket =
3µmHZvthfice
4ρpa
= 1.1 × 10−18 Z
ρpa
,
(12)
where Z is the sublimation rate of the particle (molec cm−2 s−1), µ is the molecular weight
of the sublimating ice (18 u for water), mH is the mass of hydrogen (g), and fice is the ice
fraction of the particle (all remaining parameters are in cgs units). Following Reach et al.
(2009), we define the dimensionless rocket effect parameter α as the ratio of the force from
the sublimation mass-loss, Frocket, to the force of gravity from the Sun, Fgrav:
α ≡ Frocket
Fgrav
= 2.238 × 1026 3µmHZvthfice
4GM(cid:12)ρpa
,
(13)
where G is the gravitational constant (cm3 g−1 s−2), and M(cid:12) is the mass of the Sun (g). The
rocket effect parameter should have a heliocentric distance dependency, since the product
Zvth does not necessarily vary as r−2
h , but Eq. 13 will serve as a good approximation for
small ∆rh. The α parameter is analogous to the β parameter for dust; both parameters
quantify a force directed away from the Sun in fractions of the solar gravitational force.
Assuming no losses from scattering or thermal emission, the maximum water sublima-
tion rate from ice (Zmax) can be computed from the solar flux density, particle absorption
efficiency (Qabs), and the latent heat of sublimation for water ice (L),
Zmax =
S(cid:12)QabsNA
r2
hL
(14)
where NA is Avogadro's number, and L ranges from 5.0 × 1011 to 5.1 × 1011 erg mol−1 for
temperatures from 100 to 300 K (Murphy and Koop 2005). For large compact aggregates
(i.e., porous spheres) and solid particles Qabs ≈ 1, but it is larger for fluffy aggregates since
they have light scattering properties more like a collection of monomers, rather than a single
solid particle (Kolokolova et al. 2007). This last comment aside, Zmax becomes 1.4 × 1018
molec cm−2 s−1 for T = 300 K. To account for scattering, Zmax will scale with 1 − AB,
where AB is the Bond albedo. In §6.2 we computed AB = 0.84 for our icy particle model.
Scattering reduces the Zmax of the icy model to Zmax,icy = 2.2 × 1017 molec cm−2 s−1. For
our dusty model, AB = 0.013 and Zmax,dusty ≈ Zmax.
For comparison, consider the model of Beer et al. (2006) for the sublimation rate of icy
particles. They employ spherical grains, heated by absorption of sunlight and cooled by
33
thermal emission and sublimation, and computed the absorption and emission efficiencies
with Mie scattering and effective medium theory considering both pure ice grains, and "dirty-
ice" grains, i.e., ice mixed with a generic absorber (dust). Mixing dust with the ice has a
significant effect on the ice equilibrium temperature, but the effect is not a strong function
of the ice-to-dust mass ratio (they tested mice/mdust = 0.9 and 0.5). Dust should also affect
the overall albedo of the particles, but they do not report this parameter. In their Figs. 8
and 9 they present computed grain lifetimes, defined as the time a grain takes to completely
sublimate. Their lifetimes for rh = 1.09 AU are the best examples for our scenario. For
a > 0.01 cm, pure ice grains have lifetimes of τpure = 1 × 1010a s for particle radius a
measured in cm. Dirty-ice grains have a much shorter lifetime: τdirty = 1 × 105a s. Since
the large particles spend most of their lifetime larger than 0.01 cm, we can transform their
lifetimes into sublimation rates:
4ρpa
.
Z =
3µmHτ fice
(15)
The Beer et al. (2006) model sublimation rates for large particles are Zdirty = 4 × 1017
molec cm−2 s−1, and Zpure = 4 × 1012 molec cm−2 s−1. These values are comparable to or
less than our maximum sublimation rates.
Taking Zmax,icy, a 10 cm radius particle is accelerated at a rate arocket = 0.024 cm s−2
away from the Sun, yielding a rocket parameter α = 0.042, i.e., the rocket effect is 4% the
force of solar gravity. For Z = Zmax, we compute α = 0.27. Unlike radiation pressure, the
rocket effect is potentially very strong.
Acceleration from sublimation will distribute the particles in the anti-solar direction.
Following our method for radiation pressure, we can constrain the particle ejection speed
with the implied sublimation rate
ej = 2.2 × 10−18 Zd
v2
ρpa
.
(16)
Again, adopting 2 -- 4 km as our typical turnaround distance, and taking Z = Zmax,icy, we find
ejection speeds of 310 − 440a−1/2 cm s−1 for our solid ice case (990 − 1400a−1/2 cm s−1 for
icy aggregates). These ejection speeds are higher than the instantaneous speeds measured
by Hermalyn et al. (this issue), but the two speeds do not need to agree since one is at/near
the surface and the other is out in the coma. For Z = Zmax, the resulting ejection speeds
are increased: vej = 780 − 1100a−1/2 cm s−1 (solid ice) and 2500 − 3500a−1/2 cm s−1 (icy
aggregates). However, in the above analysis we have assumed that only the sunlit hemisphere
is sublimating. By distributing the sublimation across more of the surface, we can decrease
the implied ejection speeds to 10 -- 100 cm s−1, similar to the speed measured in the coma by
Hermalyn et al. (this issue). Therefore, we conclude that a sublimation rate excess on the
sunlit hemisphere of order 1017 molec cm−2 s−1 readily describes the sunward/anti-sunward
particle asymmetry. Detailed simulations will be needed to fully account for the observed
distribution of particle velocities.
7. Mass and water production rate
Table 4 includes the total particle mass, based on our icy photometric model and a particle
density of 1 g cm−3. For the dusty model and ρp = 0.3 g cm−3, the masses are increased by
34
a factor of 672. With our preferred flux lower limit of 0.1 -- 1.0 × 10−12, the particle masses
correspond to 0.03− 0.10MN (icy), 23− 70MN (dusty), where MN = 2.4× 1011 g is the mass
of the nucleus, assuming a 0.3 g cm−3 density (Thomas et al. this issue). It is clear that the
dusty case is impossible. The only way we can reduce the estimated total mass for these
dark particles is by decreasing their densities to well below 10−3 g cm−3. Therefore, we do
not favor the dusty case, but it does remain as a possible interpretation. Given that the total
mass lost from the comet per orbit is of order 1% of the nucleus (Thomas et al. this issue),
even the solid ice particles may be too massive. Porous ice particles (e.g., ρp = 0.1 g cm−3)
should be considered the most likely case.
Icy particles will begin sublimating as soon as they are released from the nucleus and
warmed by insolation. In §6.3, we computed the rocket effect on the particles due to water
ice sublimation and concluded that a sublimation rate excess of 1017 molec cm−2 s−1 on the
particle's sunlit hemisphere can describe the observed sunward/anti-sunward asymmetry
in particle column density. We also computed the maximum sublimation rate, based on
energy balance between absorbed solar radiation and sublimation. We apply this latter
value, Zmax,icy to all of the large particles and compute water production rates. The total
large particle water production rate within a 20.6 km radius aperture is limited to < 0.6 −
5 × 1025 molec s−1, which is < 0.1 − 0.5% of the total water production rate of the comet
(≈ 1 × 1028 molec s−1). We can change the water production rate by assuming a different
photometric model as Q ∝ (1 − AB)(ApΦ)−1. For our nucleus-like case, we find Q(H2O) <
16 − 80% of the total water production rate, but, unless the particles are fluffy aggregates
with ρ (cid:46) 10−3 g cm−3, we rule out this case based on their mass. The icy particle case yields
our best estimate of the water production rate, Q(H2O) < (1 − 5 × 10−3)Qtotal.
8. Comparison to other observations
Harmon et al. (2011) observed comet Hartley 2 with Arecibo S-band (λ = 12.6 cm)
radar at the end of October 2010, about a week before Deep Impact's closest-approach. In
their average Doppler spectrum, they observe a strong grain-coma echo, with a characteristic
radial velocity dispersion of 4 m s−1. The velocity distribution is asymmetric, with a range
of ≈ −50 to +13 m s−1 with respect to the nucleus (negative velocities are away from the
Earth). The coma has a strongly depolarized echo, which indicates the radii of the largest
particles are well above the Rayleigh limit of λ/2π = 2 cm. Harmon et al. (2011) suggest that
there exists a significant population of particles with decimeter sizes or larger. We explore
the possibility that the radar observations may be the large particles imaged by Deep Impact.
The average radar cross section of the coma was 0.89 km2. From our MRI observations,
we derived a total icy particle cross section of σ = 4 × 10−4 to 3 × 10−3 km2 within 20.6 km
from the nucleus. Our cross section is more than two orders of magnitude smaller than
the radar observed cross section. However, their beam is much larger than what the MRI
can image when individual particles are detectable. We have found particles out to 40 km
in Fig. 13, but the Arecibo beam size is 32,000 km at the distance of the comet. With
speeds of order 4 m s−1, and lifetimes of order 105 s (dirty ice), the large icy particle coma
would extend out to ∼ 400 km. This estimate implies our census of the large particles is
incomplete by about a factor of < 400/20.9 = 20, resulting in a total icy particle cross section
of (cid:46) 0.01 km2, which still remains inconsistent with the radar results.
35
Assuming for the moment that the radar cross section properly reflects the total icy
particle population we compute an upper limit to the total water production rate of <
2 × 1027 molec s−1. The SWAN instrument on the SOHO satellite observes Lyα emission
over fields of view much larger than the radar beam size (105 km pixel−1). These observations
are a good point of reference for a total water production rate that is sure to include water
produced by any large particles observed with Arecibo. Combi et al. (2011) measured a water
production rate of 6 -- 9×1027 molec s−1 near perihelion (Combi et al. 2011). By this estimate,
it seems that the large particles could account for a substantial fraction of the total water
production rate of the comet. Note, however, that the SWAN-based water production rates
are on par with those observed in 4000 km apertures and smaller. Based on the aperture
sizes and water production rates listed in Table 5, most of the water is produced close to the
nucleus, perhaps within a few tens or hundreds of kilometers, suggesting that the Arecibo
observed particles have a low water sublimation rate, if any.
A dusty large particle population yields a cross section of σ = 0.07− 0.5 km2 (§5), which
is in better agreement with the radar results, but suggests that the entire large particle coma
is within ∼ 40−−250 km from the comet. There is no indication in Figs. 14 and 17 that the
large particle coma is truncated on this length scale. Moreover, dusty particles may not have
sublimating ice. Instead, they could fragment into finer particles. If the large particles are
dusty, they must fragment on (cid:46) 40−−250 km length scales in order to keep their total cross
section less than or equal to the observed radar cross section. Just based on the observed
cross sections, we consider the dusty case to be less likely than the icy case, but still find the
icy case to be lacking. Of course, a coma of both icy and dusty particles is possible. More
information on the composition and light scattering properties, including radar wavelengths,
will be needed to reconcile the radar and MRI observations.
9. Summary
Comet Hartley 2 is surrounded by a coma of large particles with radii (cid:38) 1 cm. With
observations from Deep Impact, we measured their total flux and flux distribution, based
on photometry of individual particles. The flux distribution of these particles implies a
very steep size distribution with power-law slopes ranging from −6.6 to −4.7. We estimate
that the particles account for 2 -- 14% of the total flux from the near-nucleus coma. The
spatial distribution of the particles is biased to the anti-sunward direction, as observed by
the spacecraft both pre- and post-closest approach. Radial expansion from the active areas of
the rotating nucleus does not explain the observed spatial distribution, even if the ejection
speeds are very low (∼ 1 cm s−1). Radiation pressure from sunlight cannot redistribute
them into the anti-sunward direction on small enough length scales unless the particles have
extremely low densities (∼ 10−4 g cm−3) or low radial ejection velocities ((cid:46) 10 cm s−1). Low
ejection velocities suggest there should be a strong anti-sunward velocity component in the
coma, but this does not agree with the velocity distribution observed by Hermalyn et al.
(this issue).
We examined two possible particle compositions. Our models were based on the photo-
metric properties of the nucleus of Hartley 2 (dusty case: low albedo, 0.3 g cm−3) and the
Jovian satellite Europa (icy case: high albedo, 0.1 -- 1.0 g cm−3), and serve as approximate
limiting cases.
36
The dusty case produces particle size estimates ranging from 10 cm to 2 m in radius,
the largest of which is just at the limiting resolution of Deep Impact's HRIVIS camera at
closest approach. Such large particles may be lifted off the nucleus by gas drag if CO2 is the
driving gas. Water is a plausible alternative if the water production rate from the nucleus
is at least 1027 molec s−1. Based on the dusty model, the total large particle cross section
within 20.6 km from the nucleus is 0.07-0.5 km2, similar to the 0.89 km2 radar cross section
observed by Harmon et al. (2011). If these particles are mini-nuclei, we estimate they account
for 16 -- 80% of the comet's total water production rate (within 20.6 km). However, we can
all but rule out the dusty case based on total mass estimates of the large particles, which
are in excess of 10 nucleus masses for densities of 0.3 g cm−3. Densities (cid:46) 10−3 g cm−3 are
required to reduce the total mass of the particles to a few percent of the nucleus, which is
needed to keep the particle mass less than the total mass lost from the comet per orbit (2%
during the 2010 apparition; Thomas et al. this issue).
The icy case produces particle size estimates ranging from 1 to 20 cm in radius. These
particles are easily lifted by water and CO2 gas drag. Icy particles would sublimate as soon as
they are heated by sunlight. If the particles have a net sublimation on their sunlit sides, they
would feel a rocket force that could easily distribute the particles into the anti-sunward direc-
tion. The sublimation rate excess required for this redistribution is (cid:46) 1017 molec cm−2 s−1,
where the exact value depends on the particle ejection velocities. The cross section of icy
particles within 20.6 km is much smaller than the observed radar cross section by two to
three orders of magnitude. The water production rate of the Deep Impact observed particles
is limited to < 0.1 − 0.5% of the comet's total water production rate. The total mass of
the particles for a density of 1 g cm−3 is 3 -- 10% the total mass of the nucleus. Thus, porous
aggregates with ρp (cid:46) 0.1 g cm−3 should be considered more likely than solid ice particles.
We consider the icy case to be more likely than the dusty case for three reasons: (1) the
icy particles are more easily lifted by gas drag; (2) we can account for the sunward/anti-
sunward asymmetry in the particle distribution if ice is sublimating on their sunlit sides;
and (3) the total large particle mass for the dusty case is much greater than the total mass
of the nucleus. However, several details are needed in order to test our hypothesis. We need
an improved icy particle model that treats the large particles more like macroscopic objects
to better understand their water production rates (if icy) and light scattering properties
(dusty or icy). We need to better constrain the particle densities, which may rule out the
dusty case. A hydrodynamic analysis of the near-nucleus coma, especially around the highly
active small end, would improve our knowledge of the dynamics of large particles. A better
understanding of the radar coma that includes grains smaller than λ/2π could help resolve
the discrepancy between our icy particle cross section and the observed radar coma cross
section.
If indeed icy with a high albedo, the large particles do not appear to be the source of the
comet's enhanced water production rate; although, as discussed above, there is much work
that can be done to refine this conclusion. We suspect that the small icy grains in the jets,
as observed in Deep Impact IR spectra (A'Hearn et al. 2011, Protopapa et al. 2011), are a
significant source of water, and the primary cause of the hyperactivity of Comet Hartley 2.
37
Acknowledgments
The authors thank Lev Nagdimunov (UMD) for assistance in computing cluster aggregate
porosities, Adam Ginsberg for providing the power-law fitting code, and Bjorn Davidsson
and an anonymous referee for helpful comments that improved this manuscript.
This work was supported by NASA's Discovery Program contract NNM07AA99C to the
University of Maryland and task order NMO711002 to the Jet Propulsion Laboratory.
This research made use of the PyRAF and PyFITS software packages available at http:
//www.stsci.edu/resources/software_hardware. PyRAF and PyFITS are products of
the Space Telescope Science Institute, which is operated by AURA for NASA.
References
A'Hearn, M. F., et al., 2011. EPOXI at Comet Hartley 2. Science 332, 1396 -- 1400.
Barry, R. K., et al., 2010. Development and utilization of a point spread function for the
Extrasolar Planet Observation and Characterization/Deep Impact Extended Investigation
(EPOXI) Mission. In: Society of Photo-Optical Instrumentation Engineers (SPIE) Con-
ference Series. Vol. 7731 of Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference Series.
Bauer, J. M., et al., 2011. WISE/NEOWISE Observations of Comet 103P/Hartley 2. Astro-
phys. J. 738, 171.
Beer, E. H., Podolak, M., Prialnik, D., 2006. The contribution of icy grains to the activity
of comets. I. Grain lifetime and distribution. Icarus 180, 473 -- 486.
Belton, M. J. S., et al., this issue. The complex spin state of 103P/Hartley 2. Kinematics
and orientation in space. Icarus.
Bonev, B. P., et al., this issue. Evidence for Two Modes of Water Release in Comet
103P/Hartley 2: Distributions of Column Density, Rotational Temperature, and Ortho-
Para Ratio. Icarus.
Buratti, B., Veverka, J., 1983. Voyager photometry of Europa. Icarus 55, 93 -- 110.
Burns, J. A., Lamy, P. L., Soter, S., 1979. Radiation forces on small particles in the solar
system. Icarus 40, 1 -- 48.
Clauset, A., Shalizi, C. R., Newman, M. E. J., 2009. Power-law distributions in empirical
data. SIAM Review 51 (4), 661 -- 703.
URL http://link.aip.org/link/?SIR/51/661/1
Combi, M. R., Bertaux, J.-L., Qu´emerais, E., Ferron, S., Makinen, J. T. T., 2011. Water
Production by Comet 103P/Hartley 2 Observed with the SWAN Instrument on the SOHO
Spacecraft. Astrophys. J., Lett. 734, L6.
Cowan, J. J., A'Hearn, M. F., 1979. Vaporization of comet nuclei - Light curves and life
times. Moon Planet. 21, 155 -- 171.
38
Dello Russo, N., et al., 2011. The Volatile Composition and Activity of Comet 103P/Hartley
2 During the EPOXI Closest Approach. Astrophys. J., Lett. 734, L8.
Epifani, E., et al., 2001. ISOCAM Imaging of Comets 103P/Hartley 2 and 2P/Encke. Icarus
149, 339 -- 350.
Fulle, M., 2004. Motion of cometary dust. In: Festou, M. C., Keller, H. U., Weaver, H. A.
(Eds.), Comets II. The University of Arizona Press, Tucson, AZ, pp. 565 -- 575.
Green, S. F., et al., 2004. The dust mass distribution of comet 81P/Wild 2. J. Geophys. Res.
(Planet.) 109 (E18), 12.
Groussin, O., Lamy, P., Jorda, L., Toth, I., 2004. The nuclei of comets 126P/IRAS and
103P/Hartley 2. Astron. Astrophys. 419, 375 -- 383.
Grundy, W. M., et al., 2007. New Horizons Mapping of Europa and Ganymede. Science 318,
234 -- 237.
Hampton, D. L., et al., 2005. An Overview of the Instrument Suite for the Deep Impact
Mission. Space Sci. Rev. 117, 43 -- 93.
Hanner, M. S., Giese, R. H., Weiss, K., Zerull, R., 1981. On the definition of albedo and
application to irregular particles. Astron. Astrophys. 104, 42 -- 46.
Harker, D. E., Wooden, D. H., Woodward, C. E., Lisse, C. M., 2002. Grain Properties of
Comet C/1995 O1 (Hale-Bopp). Astrophys. J. 580, 579 -- 597.
Harker, D. E., et al., 2011. Mid-infrared Spectrophotometric Observations of Fragments B
and C of Comet 73P/Schwassmann-Wachmann 3. Astron. J. 141, 26.
Harmon, J. K., Nolan, M. C., Howell, E. S., Giorgini, J. D., Taylor, P. A., 2011. Radar
Observations of Comet 103P/Hartley 2. Astrophys. J., Lett. 734, L2.
Harmon, J. K., Nolan, M. C., Ostro, S. J., Campbell, D. B., 2004. Radar studies of comet
nuclei and grain comae. In: Festou, M. C. and Keller, H. U. and Weaver, H. A. (Ed.),
Comets II. The University of Arizona Press, Tucson, pp. 265 -- 279.
Hermalyn, B., et al., this issue. The detection, localization, and dynamics of Large Icy
Particles Surrounding 103P/Hartley 2. Icarus.
Ishiguro, M., et al., 2002. First Detection of an Optical Dust Trail along the Orbit of
22P/Kopff. Astrophys. J., Lett. 572, L117 -- L120.
Kelley, M. S., Reach, W. T., Lien, D. J., 2008. The dust trail of Comet 67P/Churyumov-
Gerasimenko. Icarus 193, 572 -- 587.
Klaasen, K. P., et al., 2008. Invited Article: Deep Impact instrument calibration. Rev. Sci.
Instrum. 79 (9), 091301.
Klaasen, K. P., et al., in prep. EPOXI Instrument Calibration.
39
Knight, M. M., Schleicher, D. G., this issue. The highly unusual outgassing of Comet
103P/Hartley 2 from narrowband photometry and imaging of the coma. Icarus.
Kolokolova, L., Hanner, M. S., Levasseur-Regourd, A., Gustafson, B. A. S., 2004. Physical
properties of cometary dust from light scattering and thermal emission. In: Festou, M. C.,
Keller, H. U., Weaver, H. A. (Eds.), Comets II. The University of Arizona Press, Tucson,
pp. 577 -- 604.
Kolokolova, L., Kimura, H., Kiselev, N., Rosenbush, V., 2007. Two different evolutionary
types of comets proved by polarimetric and infrared properties of their dust. Astron.
Astrophys. 463, 1189 -- 1196.
Li, J.-Y., et al., this issue. Photometry of the nucleus of Comet 103P/Hartley 2. Icarus.
Lindler, D., Busko, I., A'Hearn, M. F., White, R. L., 2007. Restoration of Images of Comet
9P/Tempel 1 Taken with the Deep Impact High Resolution Instrument. Publ. Astron.
Soc. Pac. 119, 427 -- 436.
Lindler, D., Heap, S., Holbrook, J., Malumuth, E., Norman, D., Vener-Saavedra, P. C., 1994.
Star Detection, Astrometry, and Photometry in Restored PC Images. In: R. J. Hanisch
& R. L. White (Ed.), The Restoration of HST Images and Spectra - II. Space Telescope
Science Institute, Baltimore, pp. 286 -- 295.
Lindler, D. J., A'Hearn, M. F., Besse, S., Klaasen, K. P., this issue. Interpretation of Results
of Deconvolved Images from the Deep Impact Spacecraft High Resolution Instrument.
Icarus.
Lisse, C. M., et al., 1998. Infrared Observations of Comets by COBE. Astrophys. J. 496,
971 -- 991.
Lisse, C. M., et al., 2009. Spitzer Space Telescope Observations of the Nucleus of Comet
103P/Hartley 2. Publ. Astron. Soc. Pac. 121, 968 -- 975.
McDonnell, J. A. M., et al., 1987. The dust distribution within the inner coma of comet
P/Halley 1982i - Encounter by Giotto's impact detectors. Astron. Astrophys. 187, 719 --
741.
McLaughlin, S. A., Carcich, B., Sackett, S., Klaasen, K. P., 2011a. EPOXI 103P/Hartley 2
Encounter - HRIV Calibrated Images V1.0, DIF-C-HRIV-3/4-EPOXI-HARTLEY2-V1.0.
NASA Planet. Data Syst.
McLaughlin, S. A., Carcich, B., Sackett, S., Klaasen, K. P., 2011b. EPOXI 103P/Hartley
2 Encounter - MRI Calibrated Images V1.0, DIF-C-MRI-3/4-EPOXI-HARTLEY2-V1.0.
NASA Planet Data Syst.
Meakin, P., 1984. Effects of cluster trajectories on cluster-cluster aggregation: A comparison
of linear and Brownian trajectories in two- and three-dimensional simulations. Phys. Rev.
A 29, 997 -- 999.
40
Meech, K. J., et al., 2011. EPOXI: Comet 103P/Hartley 2 Observations from a Worldwide
Campaign. Astrophys. J., Lett. 734, L1.
Meech, K. J., Svoren, J., 2004. Using cometary activity to trace the physical and chemical
evolution of cometary nuclei. In: Festou, M. C., Keller, H. U., Weaver, H. A. (Eds.),
Comets II. The University of Arizona Press, Tucson, pp. 317 -- 335.
Moffat, A. F. J., 1969. A Theoretical Investigation of Focal Stellar Images in the Photo-
graphic Emulsion and Application to Photographic Photometry. Astron. Astrophys. 3,
455 -- 461.
Mumma, M. J., et al., 2011. Temporal and Spatial Aspects of Gas Release During the 2010
Apparition of Comet 103P/Hartley 2. Astrophys. J., Lett. 734, L7.
Murphy, D. M., Koop, T., 2005. Review of the vapour pressures of ice and supercooled water
for atmospheric applications. Q. J. R. Meteorol. Soc. 131, 1539 -- 1565.
Nakamura, R., Hidaka, Y., 1998. Free molecular gas drag on fluffy aggregates. Astron. As-
trophys. 340, 329 -- 334.
Protopapa, S., et al., 2011. Size distribution of icy grains in the coma of 103P/Hartley 2. In:
EPSC-DPS Joint Meeting 2011. p. 585.
Reach, W. T., Kelley, M. S., Sykes, M. V., 2007. A survey of debris trails from short-period
comets. Icarus 191, 298 -- 322.
Reach, W. T., Vaubaillon, J., Kelley, M. S., Lisse, C. M., Sykes, M. V., 2009. Distribution and
properties of fragments and debris from the split Comet 73P/Schwassmann-Wachmann 3
as revealed by Spitzer Space Telescope. Icarus 203, 571 -- 588.
Snyder, D. L., Hammoud, A. M., White, R. L., 1993. Image recovery from data acquired
with a charge-coupled-device camera. J. Opt. Soc. Am. A 10, 1014 -- 1023.
Sykes, M. V., Walker, R. G., 1992. Cometary dust trails. I - Survey. Icarus 95, 180 -- 210.
Thomas, P. C., et al., this issue. Shape, density, and geology of the nucleus of Comet
103P/Hartley 2. Icarus.
Tody, D., 1993. IRAF in the Nineties. In: R. J. Hanisch, R. J. V. Brissenden, & J. Barnes
(Ed.), Astronomical Data Analysis Software and Systems II. Vol. 52 of Astronomical
Society of the Pacific Conference Series. p. 173.
van de Hulst, H. C., 1957. Light Scattering by Small Particles. John Wiley and Sons, New
York.
Vaubaillon, J. J., Reach, W. T., 2010. Spitzer Space Telescope Observations and the Particle
Size Distribution of Comet 73P/Schwassmann-Wachmann 3. Astron. J. 139, 1491 -- 1498.
41
Table 1: HRIVIS and MRI images considered in this paper. Table columns are: t − tenc, time relative to
encounter; Exp., exposure time; ∆, spacecraft-comet distance; Scale, pixel scale at the distance of the comet;
φ, phase (Sun-comet-spacecraft) angle.
t − tenc
(s)
−48.4
−39.4
−29.1
−18.8
−9.8
−0.7
9.6
19.8
30.2
40.1
−414.0
−326.5
−254.4
−196.8
−167.0
−139.8
−119.9
−109.1
−100.3
−89.5
−79.6
−69.9
−59.0
−54.2
−49.2
−44.5
−39.5
−34.7
−29.9
−25.0
−20.0
−15.3
−10.3
−5.5
−0.6
4.2
9.1
14.0
18.9
23.7
28.6
33.5
38.5
44.2
49.1
54.1
59.0
68.9
78.8
88.7
98.7
109.5
119.5
142.4
171.1
207.1
265.7
335.2
403.9
Image
hv5004024
hv5004025
hv5004027
hv5004028
hv5004030
hv5004031
hv5004033
hv5004034
hv5004036
hv5004037
mv5002051
mv5002061
mv5002069
mv5004001
mv5004005
mv5004009
mv5004012
mv5004014
mv6000000
mv5004021
mv5004023
mv5004025
mv5004027
mv5004029
mv5004030
mv5004031
mv5004032
mv6000001
mv5004040
mv5004041
mv5004042
mv5004044
mv5004045
mv5004046
mv6000002
mv5004051
mv5004052
mv5004053
mv5004054
mv5004056
mv5004057
mv5004058
mv6000003
mv5004061
mv5004062
mv5004063
mv5004064
mv5004066
mv5006000
mv5006001
mv6000004
mv5006011
mv5006012
mv5006016
mv5006020
mv5006024
mv5006030
mv5006037
mv5006046
Exp.
(ms)
∆ Scale
(m)
(km)
φ
(◦)
HRIVIS
1.8
1.7
1.6
1.5
1.4
1.4
1.4
1.5
1.6
1.7
51.5
40.8
32.1
25.2
21.7
18.6
16.3
15.1
14.2
13.0
12.0
11.1
10.1
9.6
9.2
8.8
8.5
8.2
7.9
7.6
7.4
7.2
7.1
7.0
6.9
7.0
7.0
7.2
7.3
7.5
7.8
8.1
8.4
8.8
9.2
9.6
10.0
11.0
11.9
12.9
14.0
15.2
16.3
18.9
22.2
26.4
33.5
41.9
50.2
79.6
79.3
79.0
79.0
79.1
79.6
80.5
81.5
82.7
83.8
84.8
84.4
84.0
83.4
83.0
82.6
82.1
81.8
81.6
81.2
80.9
80.5
80.1
79.9
79.7
79.5
79.3
79.2
79.0
79.0
79.0
79.0
79.1
79.3
79.6
80.0
80.4
80.9
81.4
82.0
82.6
83.1
83.7
84.3
84.8
85.2
85.7
86.5
87.2
87.7
88.2
88.7
89.1
89.8
90.4
91.0
91.6
92.1
92.4
915
847
781
732
704
694
704
735
787
851
5148
4082
3210
2523
2172
1858
1632
1513
1418
1303
1202
1107
1006
964
922
885
848
815
786
760
737
719
706
697
694
696
703
715
732
753
778
807
840
882
921
962
1004
1096
1192
1294
1399
1517
1627
1886
2219
2643
3345
4187
5023
375.5
125.5
375.5
125.5
375.5
125.5
375.5
125.5
375.5
125.5
MRI
500.5
500.5
500.5
500.5
500.5
500.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
120.5
40.5
500.5
500.5
500.5
500.5
500.5
500.5
42
Table 2: Online only table. All single power-law flux distributions derived from HRIVIS and MRI
images. Table columns: Fλ, the minimum, maximum, and total particle fluxes in the fit; N , number
of particles in the fit; α, best-fit power-law slope and uncertainty; PKS, the Kolmogorov-Smirnov
probability that the distribution is a power-law; FT /Fcoma, the fraction of the total coma flux
attributable to point sources with Fλ > Fλ,min in MRI full-frame data, excluding regions near the
nucleus.
Image
∆
(km)
a
Fλ
min max
(10−12 W m−2 µm−1)
total
N
α
PKS
FT /Fcoma
hv5004024
hv5004025
hv5004027
hv5004028
hv5004030
hv5004031
hv5004033
hv5004034
hv5004036
hv5004037
mv5004029
mv5004031
mv6000001
mv5004041
mv5004044
mv5004046
mv5004051
mv5004053
mv5004056
mv5004058
mv5004027
mv5004030
mv5004032
mv5004040
mv5004042
mv5004045
mv6000002
mv5004052
mv5004054
mv5004057
mv6000003
mv5004062
mv5004064
mv5004021
mv5004025
mv5004029
mv5004031
915
847
781
732
704
694
704
735
787
851
964
885
815
760
719
697
696
715
753
807
1006
922
848
786
737
706
694
703
732
778
840
921
1004
1303
1107
964
885
1.8
1.5
1.4
1.5
0.9
1.5
0.9
1.5
0.9
1.5
4.1
4.5
4.2
3.9
3.8
3.6
3.0
3.1
2.8
3.0
5.5
6.7
6.0
6.1
6.0
6.2
5.9
5.8
5.7
5.9
5.6
5.3
4.9
2.4
2.8
3.4
3.5
20.5
31.6
29.0
33.1
35.5
40.6
34.3
31.2
23.1
21.2
14.1
13.4
11.4
16.6
18.8
12.9
13.8
21.3
11.1
9.0
16.4
17.8
22.6
24.3
24.1
22.2
21.1
26.2
27.1
21.4
19.8
18.3
15.7
10.6
14.2
13.9
16.6
HRIVIS
1853
3128
1388
2058
1222
1252
918
1017
902
1005
(%)
641
986
430
685
613
419
379
374
496
425
-3.06± 0.09
0.3
-2.88± 0.06 <0.1
-2.86± 0.08
0.6
-2.91± 0.07
12.9
-2.67± 0.07
16.8
-2.70± 0.09
13.4
-2.67± 0.08
15.2
-2.93± 0.10
15.3
-2.94± 0.09
8.6
-3.08± 0.11
41.7
112
114
124
135
141
124
155
110
135
116
133
108
189
252
321
331
390
374
309
208
235
188
149
MRI sub-frame
692± 7
660± 7
770± 7
836± 7
933± 7
631± 5
729± 6
553± 5
637± 5
437± 4
MRI 41 ms
1066± 9
884± 9
1773± 12
2053± 13
2483± 15
2747± 15
3406± 17
3100± 16
2822± 15
1977± 13
2029± 13
1359± 10
922± 9
MRI 121 ms
2993± 13
4374± 16
6086± 20
7945± 23
838
1191
1291
1501
-4.08± 0.29
-4.31± 0.31
-4.03± 0.27
-3.87± 0.25
-3.44± 0.21
-3.50± 0.22
-3.22± 0.18
-3.36± 0.22
-3.34± 0.20
-3.83± 0.26
-4.05± 0.26
-4.39± 0.33
-3.76± 0.20
-4.04± 0.19
-4.02± 0.17
-3.96± 0.16
-3.68± 0.14
-3.74± 0.14
-3.59± 0.15
-3.60± 0.18
-3.78± 0.18
-4.08± 0.22
-3.96± 0.24
38.5
10.1
8.2
54.5
3.0
36.4
13.5
99.1
35.7
41.9
67.3
74.5
29.8
97.8
43.0
29.9
0.3
8.7
30.0
7.3
29.4
47.0
33.8
-3.95± 0.10
8.4
-3.87± 0.08
3.2
-3.88± 0.08
0.6
-3.87± 0.07 <0.1
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
···
0.0007
0.0006
0.0012
0.0019
0.0032
0.0040
0.0046
0.0044
0.0035
0.0023
0.0018
0.0011
0.0008
0.0021
0.0031
0.0039
0.0048
Continued on next page
43
Image
mv6000001
mv5004041
mv5004044
mv5004046
mv5004051
mv5004053
mv5004056
mv5004058
mv5004061
mv5004063
mv5004066
mv5006001
∆
(km)
815
760
719
697
696
715
753
807
882
962
1096
1294
total
19.7
19.8
21.0
25.4
29.2
23.1
20.7
16.8
24.0
14.3
14.7
12.2
min max
(10−12 W m−2 µm−1)
7337± 22
4.1
9078± 26
4.1
9536± 26
3.8
9012± 25
4.3
9695± 26
3.7
9603± 26
3.5
9543± 26
3.4
10378± 27
3.2
8495± 24
3.0
6967± 21
2.6
4796± 17
2.4
2373± 11
2.1
MRI 501 ms
54± 1
102± 1
243± 2
653± 3
1052± 5
1526± 7
1288± 6
1028± 5
400± 3
166± 2
81± 1
0.4
0.5
0.7
0.8
0.9
1.1
0.9
0.7
0.7
0.7
0.5
1.1
1.7
3.0
3.4
5.6
6.7
7.3
4.3
3.2
2.8
1.6
mv5002051
mv5002061
mv5002069
mv5004001
mv5004005
mv5004009
mv5006016
mv5006020
mv5006024
mv5006030
mv5006037
a The absolute calibration uncertainties are 5% for HRIVIS, and 10% for MRI.
-4.54± 0.35
-4.70± 0.27
-4.26± 0.19
-3.76± 0.10
-3.70± 0.09
-3.69± 0.08
-3.59± 0.08
-3.72± 0.09
-3.81± 0.14
-4.04± 0.22
-4.85± 0.35
5148
4082
3210
2523
2172
1858
1886
2219
2643
3345
4187
105
192
292
721
837
1034
991
954
411
184
120
82.3
92.6
5.8
11.6
17.6
0.1
9.0
4.2
39.6
63.0
37.5
Continued
a
Fλ
N
α
PKS
FT /Fcoma
(%)
-3.95± 0.08
0.5
-3.91± 0.08
0.2
-3.60± 0.06 <0.1
-3.76± 0.07
0.1
-3.65± 0.06 <0.1
-3.57± 0.06 <0.1
-3.58± 0.06 <0.1
-3.65± 0.06 <0.1
-3.65± 0.06 <0.1
-3.64± 0.06
1.2
-3.82± 0.07
1.2
-3.87± 0.09
17.8
1262
1468
1770
1391
1769
1769
1806
1937
1681
1771
1426
965
0.0050
0.0068
0.0089
0.0087
0.0104
0.0101
0.0083
0.0073
0.0056
0.0047
0.0034
0.0021
0.0002
0.0003
0.0005
0.0011
0.0013
0.0017
0.0015
0.0013
0.0007
0.0003
0.0002
44
Table 3: Summary of single power-law fits to individual HRIVIS and MRI flux distributions listed in Table 2.
Table columns: Fλ, mean minimum and maximum particle fluxes; N , mean number of particles; α, minimum,
maximum, mean, and standard deviation of the flux distribution power-law slopes; PKS, mean Kolmogorov-
Smirnov probability that the distributions are power-laws; FT /Fcoma, mean fraction of the total coma flux
attributable to point sources with Fλ > Fλ,min in MRI full-frame data, excluding regions near the nucleus.
Instrument
∆
Fλ
N
α
PKS
FT /Fcoma
min max min max
min max mean
σ
(km)
(10−12
W m−2 µm−1)
HRI
MRI sub-frame
MRI 41 ms
MRI 121 ms
MRI 501 ms
MRI (distant)a
694
696
694
696
1858
2172
915 1.3
964 3.6
1006 5.5
1303 3.1
5148 0.7
2643 9.0
30.0
14.2
20.2
17.0
3.5
46.9
545
127
245
1489
531
731
-3.08
-4.31
-4.39
-3.95
-4.85
-3.81
-2.67
-3.22
-3.59
-3.57
-3.59
-3.70
-2.87
-3.70
-3.90
-3.76
-4.06
-3.75
0.14
0.35
0.22
0.14
0.43
0.04
(%)
0.125
0.341
0.384
0.021
0.330
0.182
···
···
0.0023
0.0059
0.0008
0.0011
a Averages of the the four 501 ms MRI images with ∆ > 2000 km and N > 300 particles, with which we
use to define the fraction of the coma flux attributable to large particles. For this row only, Fλ has been
corrected to a distance of ∆ = 700 km before averaging.
Table 4: Flux range (Fλ,min, Fλ,max), radius range (aicy), total number (N ), flux (Fλ,T ), fraction of coma
flux in particles (FT /Fcoma), cross section (σicy), mass (Micy), and water production rate (Q(H2O)) for the
large particles measured at ∆ = 2100 − 2600. The radius, cross section, mass, and water production rate
are computed using an icy composition with Europa's photometric parameters (at φ = 86◦) and a density
of 1.0 g cm−3. We also extrapolate these quantities down to 10−14 W m−2 µm−1, and to our total coma
measurement for a 20.6 km aperture at ∆ = 5148 km. Uncertainties are derived from the power-law slope
uncertainties (±0.02) and are symmetric in log space.
a
Fλ,min
Fλ,max
(W m−2 µm−1)
×10−12
a
aicy
(cm)
9.00 b
1.00
0.10 e
0.01
45 b
45
45
45
1.00
0.10 e
45
45
7.4
2.5
0.8
0.2
2.5
0.8
16.5
16.5
16.5
16.5
16.5
16.5
log10 N
log10 Fλ,T
a
FT
Fcoma
log10 σicy
log10 Micy
log10 Qmax
(W m−2 µm−1)
(%)
(cm2)
∆ ≈ 2400 km, full-frame (excluding masked regions)
2.86b
5.22
6.42±0.02
4.90±0.02
6.76±0.04
7.32±0.03
8.61±0.06
8.17±0.05
0.11b
1.8±0.1
13.8±1.1
98.7+12.8−11.3
∆ = 5148 km, 20.6 km aperture
-8.06b,c
-6.86±0.02
-5.97±0.03
-5.11±0.05
(g)
(s−1)
9.07
9.92±0.01
10.39±0.03
10.79±0.04
22.56
23.77±0.02
24.66±0.03
25.52±0.05
6.28±0.03
9.02±0.07
-6.69±0.02
-5.79±0.03
1.8±0.1 d
13.8±1.1 d
6.59±0.02
7.49±0.03
11.30±0.03
12.66±0.06
23.94±0.02
24.83±0.03
a Fluxes have been corrected to a distance of 700 km.
b Measured values.
c The instrumental and calibration uncertainties are 2 × 10−12 W m−2 µm−1 and 10%, respectively.
d Assumed value.
e Our preferred lower flux limit is between 0.1 and 1.0×10−12 W m−2 µm−1.
45
Table 5: Summary of comet Hartley 2 water production rates observed near perihelion, and derived in this
study.
Notes
Aper.a
(km)
9.1 × 105
19
19
Date(s)
(UT, 2010)
Q(H2O)
(1027 molec s−1)
1.6 × 104
1.1 × 106
3800
8.70± 0.38
3.45± 0.04
6.78± 0.26
(cid:46)2.
11.6± 0.7
6.38± 0.12
10.
11.45± 0.65
(cid:46)0.004
18 Oct
19 Oct
22 Oct
23 -- 31 Oct
31 Octb
1 Nov
2 Nov
4 Nov
4 Nov
a Slit half-width or aperture radius. We have assumed the 0.43(cid:48)(cid:48) slit for the Mumma et al. (2011) observations,
and a 3 pixel radius for the Combi et al. (2011) observations.
b Mean and standard deviation of 5 measurements taken on 31 Oct.
Combi et al. 2011
Mumma et al. 2011
Mumma et al. 2011
Large particles only, Harmon et al. 2011 + this work
Knight and Schleicher this issue
Combi et al. 2011
A'Hearn et al. 2011
Dello Russo et al. 2011
Large particles only, Table 4
300
22
20.6
46
Erratum to "A distribution of large particles in the coma of Comet
103P/Hartley 2": [Icarus 222, 634 -- 652 (2013)]
Michael S. P. Kelleya,*, Don J. Lindlerb, Dennis Bodewitsa, Michael F. A'Hearna, Carey M.
Lissec, Ludmilla Kolokolovaa, Jochen Kisseld,1, Brendan Hermalyne
aDepartment of Astronomy, University of Maryland, College Park, MD 20742-2421, USA
bSigma Space Corporation, 4600 Forbes Boulevard, Lanham, MD 20706, USA
cJohns Hopkins University -- Applied Physics Laboratory, 11100 Johns Hopkins Road, Laurel, MD 20723,
dMax-Planck-Institut fur Sonnensystemforschung, Max-Planck-Str. 2, 37191 Katlenburg-Lindau, Germany
eNASA Ames Research Center / SETI Institute, MS 245-3 BLDG245, Moffett Field, CA 94035
USA
1. Introduction
Comet 103P/Hartley 2 has a high water production rate to surface area ratio, suggesting
the nucleus is nearly 100% active. In contrast, images of the nucleus from the Deep Impact
Flyby spacecraft show strong localized activity, with an inner-coma enriched in CO2 gas and
water-ice grains. Rather than being produced solely by water ice sublimation at the nucleus,
the hyperactivity of Hartley 2 may be due to water-ice sublimation in the coma. However, the
contribution of coma grains is poorly constrained, leaving the icy-grain hypothesis unproven
(A'Hearn et al. 2011, Protopapa et al. 2014).
Images from the Deep Impact Flyby spacecraft show thousands of point sources surround-
ing the nucleus of comet Hartley 2 (A'Hearn et al. 2011). The point sources are individual
particles ejected by the comet. We measured their brightnesses and summarized their sizes,
total mass, and spatial distribution (Kelley et al. 2013). The photometric properties (albedo,
phase function) of the particles are unknown, therefore we adopted two models as approxi-
mate limiting cases: 1) bright, icy particles with photometric properties similar to the Jovian
satellite Europa; and 2) dark, nucleus-like particles with properties similar to the nucleus of
Hartley 2. For the bright, icy case, we reported the largest particle had a radius of 20 cm and
the total population mass within 21 km of the nucleus was up to 3 to 10% of the nucleus mass
(assuming a density of 1 g cm−3). For the dark, nucleus-like case, the largest particle was
2 m in radius and the total population up to 230 to 700% of the nucleus mass (0.3 g cm−3).
Based on this mass calculation, we ruled out the nucleus-like case and favored the icy case.
Moreover, the icy case produces particles too small to account for the comet's hyperactivity.
We have found three errors in our calculations that affected our radius and mass estimates.
We regret the errors, as they have significant consequences in our analysis of the particles.
We provide updated calculations and interpretations in Sections 2 and 3. In addition, we
provide a few minor clarifications and corrections to our original paper in Section 4.
1retired
Kelley et al. 2015, Icarus, in press
2. Radius and mass corrections
We have found an error that affects the computed particle radii (and dependent quanti-
ties), and two additional errors that affect population masses. First, Eq. 4 of Kelley et al.
(2013) is missing a factor of π in the denominator. The correct equation is
Fλ =
ApΦ(θ)σSλ,(cid:12)
πr2
h∆2
(4)
where Fλ is the particle flux density, Ap is the geometric albedo, Φ(θ) is the phase function
evaluated at the phase angle θ, σ is the cross-sectional area of the particle, Sλ,(cid:12) is the solar
flux density at 1 AU, rh is the heliocentric distance of the particle, and ∆ is the spacecraft-
√
particle distance. To account for this factor of π, all parameters derived from particle fluxes
π, cross section and water production rate
must be scaled as follows: radius by a factor of
by a factor of π, and mass by a factor of π3/2.
Second, our computations of total population mass (Table 4) erroneously used 1000 g cm−3,
rather than the 1 g cm−3 quoted in the text. In addition, our analytical formula for integrat-
ing the total mass of a population of grains given a power-law flux distribution was missing
a factor of 2. The formula was not given in the paper, but for completeness the corrected
formula is
(cid:90) amax
amin
8πρ
3
N0F α−1
1
M =
a2α+2da,
where M is the total population mass, ρ is the particle mass density, N0 is the flux distribution
normalization constant, F1 is the flux density from a 1-cm radius particle (via Eq. 4), a is the
particle radius, amin, amax are the limits of the integration (corresponding to the estimated
flux density limits), and α is the power-law index of the differential flux distribution (Eq. 3).
Thus, total population masses must be scaled by a factor of 1/500, in addition to the π3/2
scale factor from above (the total scale factor is 0.011). Other quantities that depend on
individual particle mass (e.g., β in Section 6.2, α and vej in Section 6.3) are unaffected.
We present a revised Table 4, with corrected radii, cross sections, mass, and maximum
water production rates.
In addition, all calculations are now based on the normalization
constant N0 (Eq. 6), whereas previously some calculations were normalized via the total
observed particle flux. This change increases most flux-based quantities by 13%. Overall,
our analysis is significantly affected by the revised Table 4.
The solid water-ice case (see Section 5 of Kelley et al. 2013 for definitions of our icy and
dusty cases) produces particle estimates up to 30 cm in radius with an estimated population
mass (within a 20.6 km aperture) of 0.2−6×1010 g, or up to 0.02% of the nucleus mass. This
mass is two orders of magnitude lower than the ∼ 2% of the nucleus that was lost in 2010
(Thomas et al. 2013). Porous ice particles are no longer necessary to reduce the estimated
population mass below the orbital mass loss of the comet. The total water production rate
estimate is Qmax = 0.3 − 2 × 1025 s−1, or 0.03 − 0.2% of the comet's total water production
rate, insufficient to account for the comet's hyperactivity.
To convert from the icy case to the dusty case multiply radii by 12.6, cross sections by
158, masses by 597, and Qmax by 1007. The dusty particle case produces particles up to
4 m in radius and a total population mass of 0.1 − 3 × 1013 g, or 0.6 -- 14% of the mass of the
nucleus, potentially exceeding the estimated orbital mass-loss rate of the comet. If the dusty
2
particles behave like mini-comet nuclei, their total water production rate may be as large
as Qmax = 0.3 − 2 × 1028 s−1, or 30 -- 200% of that of the comet during the encounter. Our
updated cross sections, 0.2 -- 2 km2, are comparable to and slightly exceed the cross section
measured by radar observations (0.89 km2; Harmon et al. 2011).
The new radius estimates produce particles up to 8 m diameter. Such particles should be
resolved in the High-Resolution Instrument (HRI) visible images, which has a reconstructed
resolution of approximately 3 m at closest approach. However, our original examination of
the data revealed no resolved sources, suggesting a maximum size of approximately 3 m. The
smaller size could be accommodated through an increase in the model albedo (from 0.049
to 0.31) or by reducing the phase function coefficient (from 0.046 to 0.023 mag per degree;
see Section 5 of Kelley et al. 2013). Both cases reduce the total population mass to <1% of
the nucleus mass and maximum water production rate to 40% of the total comet production
rate. We save a more thorough search for resolved HRI particles for future work.
3. Interpretation
Given the above revisions, the dusty case may solve the apparent hyperactivity of the
comet. However, if these particles are long lived (i.e., not fragmenting) and moving away
with radial velocities of order 0.1 m s−1 (Hermalyn et al. 2013), then their lifetimes in a
21 km radius aperture are of order 60 hours, implying a significantly large mass-loss rate
of ∼ 103 kg s−1 for the assumed particle density of 0.3 g cm−3. This mass-loss rate must
be increased by an order of magnitude if we instead consider the radar observed velocity
dispersion of 4 m s−1 (Harmon et al. 2011). These estimates significantly exceed the comet's
total water and carbon dioxide production rates (∼ 300 kg s−1 and ∼ 160 kg s−1, respectively;
A'Hearn et al. 2011) near the time of the flyby. Therefore, understanding the large particle
dynamics is critical to determining the erosion rate of the comet, and if the dusty case
remains valid. In parallel, a definitive upper limit to particle sizes should be derived from a
more detailed search for resolved particles in HRI images.
We have outlined two dramatically different scenarios for the physical properties of the
large particles of Hartley 2. Intermediate sets of photometric and compositional properties
are possible, and such cases cannot be ruled out. We do consider the dusty case (low albedo,
0.3 g cm−3) to be less likely due to the large and massive particles implied, but cannot
confidently rule it out at this time. However, note that higher albedos or different phase
functions may be employed to produce a more physically consistent picture for the large
particles.
4. Other corrections
Equation 5 is correctly referred to as the differential size distribution, but elsewhere the
text commonly omits the "differential" adjective. We reviewed our comparisons to other
estimates of the dust size distribution, and the nomenclature used in the literature is not
very specific, occasionally labeling differential size distributions as size distributions (just as
we did). Our best understanding of the investigations listed in the text is that most indeed
report the differential size distribution. Our conclusion, that the differential size distribution
slope, −4.7, is steeper than other estimates, remains valid.
3
The normalization constant N0 is not needed in Eq. 6, it is already included in Eq. 5.
(cid:90) Fλ,max
Fλ,min
The corrected Eq. 6 is
N =
dn
dF
dF.
(6)
The calculations for the paper used the above formula.
nucleus mass is MN = 2.4 × 1014 g (Thomas et al. 2013).
The nucleus mass listed in Section 7 was expressed in units of kg, but reported as g. The
Finally, the units of flux density in Section 4.2 should be W m−2 µm−1, and not W cm−2 µm−1
as stated.
Acknowledgments
We thank Katherine Kretke for identifying the mass discrepancy in Table 4, and Michael
Belton for motivating other clarifying comments.
This work was supported by NASA (USA) through the Planetary Mission Data Analysis
Program contract NNX12AQ64G to the University of Maryland.
References
A'Hearn, M. F., et al., 2011. EPOXI at Comet Hartley 2. Science 332, 1396 -- 1400.
Harmon, J. K., Nolan, M. C., Howell, E. S., Giorgini, J. D., Taylor, P. A., 2011. Radar
Observations of Comet 103P/Hartley 2. Astrophys. J., Lett. 734, L2.
Hermalyn, B., et al., 2013. The detection, localization, and dynamics of large icy particles
surrounding Comet 103P/Hartley 2. Icarus 222, 625 -- 633.
Kelley, M. S., et al., 2013. A distribution of
large particles in the coma of Comet
103P/Hartley 2. Icarus 222, 634 -- 652.
Protopapa, S., et al., 2014. Water ice and dust in the innermost coma of comet 103P/Hartley
2. Icarus 238, 191 -- 204.
Thomas, P. C., et al., 2013. Shape, density, and geology of the nucleus of Comet 103P/Hartley
2. Icarus 222, 550 -- 558.
4
Table 4: Flux range (Fλ,min, Fλ,max), radius range (aicy), total number (N ), total flux (Fλ,T ), fraction of
coma flux in particles (FT /Fcoma), cross section (σicy), mass (Micy), and maximum water production rate
(Qmax(H2O)) for the large particles measured at ∆ = 2100 − 2600. The radius, cross section, mass, and
water production rate are computed using an icy composition with Europa's photometric parameters (at
φ = 86◦) and a density of 1.0 g cm−3. We also extrapolate these quantities down to 10−14 W m−2 µm−1,
and to our total coma measurement for a 20.6 km aperture at ∆ = 5148 km. Uncertainties are derived from
the power-law slope uncertainties (±0.02) and are symmetric in log space.
a
Fλ,min
Fλ,max
(W m−2 µm−1)
×10−12
a
aicy
(cm)
9.00 b
1.00
0.10
0.01
1.00
0.10
45 b
45
45
45
45
45
13.1
4.4
1.4
0.4
29.3
29.3
29.3
29.3
4.4
1.4
29.3
29.3
log10 N
log10 Fλ,T
a
FT
Fcoma
log10 σicy
log10 Micy
log10 Qmax
(W m−2 µm−1)
(%)
(cm2)
∆ ≈ 2400 km, full-frame (excluding masked regions)
2.86±0.00 b
4.90±0.02
6.76±0.04
8.61±0.06
-8.01±0.00 c
-6.80±0.01
-5.91±0.03
-5.06±0.05
0.1±0.0
2.0±0.1
15.5±1.2
111±13
5.77±0.00
6.97±0.01
7.86±0.03
8.72±0.05
(g)
(s−1)
7.12±0.00
7.96±0.01
8.44±0.03
8.84±0.04
23.11±0.00
24.32±0.01
25.21±0.03
26.06±0.05
∆ = 5148 km, 20.6 km aperture
6.33±0.03
9.08±0.07
-6.63±0.01
-5.74±0.03
2.0±0.1 d
15.5±1.2 d
7.14±0.01
8.04±0.03
9.39±0.03
10.76±0.06
24.49±0.01
25.38±0.03
a Fluxes have been corrected to a distance of 700 km.
b Measured values.
c The instrumental and calibration uncertainties are 2 × 10−12 W m−2 µm−1 and 10%, respectively.
d Assumed value.
e Our preferred lower flux limit is between 0.1 and 1.0×10−12 W m−2 µm−1.
5
|
1607.03695 | 2 | 1607 | 2016-07-18T01:23:15 | Gas and dust around A-type stars at tens of Myr:signatures of cometary breakup | [
"astro-ph.EP"
] | Discs of dusty debris around main-sequence star indicate fragmentation of orbiting planetesimals, and for a few A-type stars, a gas component is also seen that may come from collisionally-released volatiles. Here we find the sixth example of a CO-hosting disc, around the 30Myr old A0-star HD 32297. Two more of these CO-hosting stars, HD 21997 and 49 Cet, have also been imaged in dust with SCUBA-2 within the SONS project. A census of 27 A-type debris hosts within 125 pc now shows 7/16 detections of carbon-bearing gas within the 5-50 Myr epoch, with no detections in 11 older systems. Such a prolonged period of high fragmentation rates corresponds quite well to the epoch when most of the Earth was assembled from planetesimal collisions. Recent models propose that collisional products can be spatially asymmetric if they originate at one location in the disc, with CO particularly exhibiting this behaviour as it can photodissociate in less than an orbital period. Of the six CO-hosting systems, only beta Pic is in clear support of this hypothesis. However, radiative transfer modelling with the ProDiMo code shows that the CO is also hard to explain in a proto-planetary disc context. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1–8 (2016)
Printed 30 June 2018
(MN LATEX style file v2.2)
Gas and dust around A-type stars at tens of Myr: signatures of
cometary break-up
J. S. Greaves1(cid:63), W.S. Holland2, B.C. Matthews3, J.P. Marshall4, W.R.F. Dent5,
P. Woitke5, M.C. Wyatt7, L. Matr`a7 & A. Jackson8
1School of Physics & Astronomy, Cardiff University, 4 The Parade, Cardiff CF24 3AA, UK
2Astronomy Technology Centre, Royal Observatory Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, UK
3National Research Council of Canada, 5071 West Saanich Rd, Victoria, BC, V9E 2E7, Canada
4School of Physics, University of New South Wales, NSW, 2052, Sydney, Australia
5ALMA Santiago Central Offices, Alonso de Crdova 3107, Vitacura, Casilla 763 0355, Santiago, Chile
6School of Physics & Astronomy, University of St Andrews, North Haugh, St Andrews, Fife KY16 9SS, UK
7Institute of Astronomy, Madingley Road, Cambridge CB3 0HA, UK
8School of Earth and Space Exploration, Arizona State University, Tempe, AZ 85287, USA
Accepted 2016. Received 2016; in original form 2016
ABSTRACT
Discs of dusty debris around main-sequence stars indicate fragmentation of orbiting plan-
etesimals, and for a few A-type stars, a gas component is also seen that may come from
collisionally-released volatiles. Here we find the sixth example of a CO-hosting disc, around
the ∼30 Myr-old A0-star HD 32997. Two more of these CO-hosting stars, HD 21997 and 49
Cet, have also been imaged in dust with SCUBA-2 within the SONS project. A census of 27
A-type debris hosts within 125 pc now shows 7/16 detections of carbon-bearing gas within the
5-50 Myr epoch, with no detections in 11 older systems. Such a prolonged period of high frag-
mentation rates corresponds quite well to the epoch when most of the Earth was assembled
from planetesimal collisions. Recent models propose that collisional products can be spatially
asymmetric if they originate at one location in the disc, with CO particularly exhibiting this
behaviour as it can photodissociate in less than an orbital period. Of the six CO-hosting sys-
tems, only β Pic is in clear support of this hypothesis. However, radiative transfer modelling
with the ProDiMo code shows that the CO is also hard to explain in a proto-planetary disc
context.
Key words: planetary systems – circumstellar matter – infrared: stars
1
INTRODUCTION
Dusty debris around main-sequence stars results from collisions be-
tween rocky bodies. Timescales for the particles to fall into the
star or grind down to sizes small enough to be blown out by ra-
diation pressure are short compared to stellar lifetimes, so larger
progenitor planetesimals must be present. In most cases, these are
found to be in belts located at tens of AU, where rock/ice comet-
like compositions are probable, akin to solar system cometary ma-
terial. Although collisions should cause the ices to sublimate into
gases, this component in the belts is difficult to detect, as molecules
are quickly photo-dissociated. For nearby debris-hosting stars, e.g.
Fomalhaut, improved limits from deep observations are important
for comparing the chemistry to solar system comets (Matr`a et al.
2015).
There are five A-type debris hosts where the molecule car-
(cid:63) E-mail: GreavesJ1 at cardiff.ac.uk
c(cid:13) 2016 RAS
bon monoxide (CO) has been detected, via millimetre rotational
transitions. This species photo-dissociates on hundreds of year
timescales even under interstellar radiation (Zuckerman & Song
2012, e.g.). When illuminated by A-stars, this timescale can be
shortened to less than orbital periods at tens of AU (Jackson et al.
2014, e.g.). If CO is preferentially produced at one location in the
disk then the CO distribution would be 'one-sided' around the star,
and this effect has been imaged recently in an ALMA study of β
Pic (Dent et al. 2014; Matr`a et al., in prep.).
Other gas phases can also be present, with CII and OI lines
in the far-infrared (Dent et al. 2012) potentially tracing photo-
dissociated CO (Roberge et al. 2013). Three of the CO-hosting
systems also appear to host 'falling evaporating bodies', with tran-
sient red-shifted Ca-absorption features seen towards β Pic and 49
Cet (Montgomery & Welsh 2012), and Na absorption identified
towards HD 32297 (Redfield 2007). These features are consistent
with volatiles originating from ongoing cometary breakups, an idea
which is now being explored by models (Kral et al. 2016, e.g.).
2
J. S. Greaves et al.
Figure 1. SCUBA-2 results for 49 Cet, on signal-to-noise ratio scales. The
peak flux at 450 microns is is 74 ± 14 mJy/beam, and the integrated flux
within a 40 arcsec diameter aperture is 125 ± 10 mJy. The dashed contours
show the 850 µm SNR (peak = 8.3), overlaid on the 450 µm colour-scale.
The integrated flux at 850 µm is 12.1 ± 2.0 mJy. The secondary peak ad-
jacent to the north (top) side of the disc may be due to a background dusty
galaxy. The stellar position coincides with the 450 µm flux-peak within typ-
ical pointing drifts ( <∼ 2(cid:48)(cid:48)).
Here we consider debris-hosting A-stars within 125 pc (par-
allax ≥8 mas) that have been searched for CO. This distance limit
helps to exclude stars outside the Local Bubble where interstellar
CO may be a strong contaminant. We report the sixth detection of
CO, around the approximately 30 Myr-old A0 star HD 32297. In
this case, we successfully used the presence of weak CII emission
(Donaldson et al. 2013) as a predictor for the presence of CO.
We have also followed CO detections for 49 Ceti and HD
21997 (Zuckerman et al. 1995; Mo´or et al. 2011) with dust
continuum-imaging at 450 and 850 micron wavelength. These con-
tinuum data are part of the JCMT Legacy Project SONS (SCUBA-2
Survey of Nearby Stars), described by Pani´c et al. (2013). The gas-
plus-dust systems are also tested here against model predictions
that the short-lived CO component could be more spatially asym-
metric than the dust.
2 OBSERVATIONS
The new data were obtained with the 15 m James Clerk Maxwell
Telescope located on Mauna Kea, Hawaii. Observing procedures
have been described by, for example, Pani´c et al. (2013, 2010).
HD 21997 and 49 Cet were observed with the SCUBA-2 cam-
era (Holland et al. 2013) in 2012-2015, in moderately dry con-
ditions (225 GHz zenith opacities under 0.1). Images are shown
with 1 arcsec pixels, and after smoothing with a 7-arcsec Gaus-
Figure 2. SCUBA-2 850 micron image for HD 21997, with contours at
3,4,5,6,7 sigma levels and a peak of 7.9 ± 1.1 mJy/beam. The integrated
flux is 10.7 ± 1.5 mJy.
sian, the effective beam diameters are 15.8 and 11.6 arcsec at 850
and 450 microns respectively. The continuum data reduction used
the makemap task in the SMURF package (Jenness et al. 2011),
plus high-pass filtering and zero-masking to remove residual low-
frequency structure in the backgrounds.
The CO data for HD 32297 at 1.3 mm were taken as part of a
poor weather backup programme in 2013 (225 GHz zenith opaci-
ties up to 0.3), using the RxA3 receiver to search for the J=2–1 tran-
sition. The spectral reduction used the makecube task in SMURF
plus the SPLAT package for baselining and binning. A 200 km/s
velocity range is shown around the stellar velocity in the heliocen-
tric frame of 23.0 ± 0.3 km/s (Torres et al. 2006). After about 10
hours on sky, the JCMT J=2-1 data are about twice as sensitive as
the corresponding spectrum of Zuckerman & Song (2012) from the
IRAM 30m telescope.
3 RESULTS
3.1 Dust imaging
Figure 1 shows the SCUBA-2 data for 49 Cet. The disc is spatially
resolved at 450 µm, at the expected orientation. For comparison,
the position angle with Herschel at 70 µm (Roberge et al. 2013) is
≈ 105◦ (anti-clockwise from north), while CO lies at PA = 110 ±
10◦ (Hughes et al. 2008). The SExtractor tool for automatic object
detection1 was used here, indicating PA ≈ 120◦ for the 450 µm disc
axis.
1 The peak just to the north was assumed to be unrelated, and was blanked
before fitting the disc.
c(cid:13) 2016 RAS, MNRAS 000, 1–8
Gas and dust around A-stars at tens of Myr
3
aperture from ALMA. Only 15 ALMA antennas were available to
Mo´or et al., so some spatial scales may not have been fully sam-
pled.
SCUBA-2 also simultaneously observed HD 21997 at 450 µm,
and a tentative signal of 85 ± 25 mJy/beam was seen near the star.
There is a similar peak about 15 arcsec SW of the star, but this
appears to coincide with a faint part of the 850 µm background
structure.
HD 32297 is not a SONS target, but has been imaged in dust
emission at 1.3 mm by Manness et al. (2008), and was resolved at
arcsec resolution. The disc was found to be extended and elliptical,
with a strong asymmetry, in that the centroid is offset from the star
with 4-sigma confidence.
3.2 CO spectroscopy
Figure 3 shows the CO spectrum towards HD 32297, reduced by
two different methods to test for robustness. The entire passband
was first fitted with a high-order polynomial, in order to fully use
the information present, and excluding only a region of 20 chan-
nels of 1.27 km/s around the stellar velocity (+23.0 km/s, Torres
et al. 2006). Secondly, a least-parameters approach was used, fit-
ting a low-order polynomial only across the two green-shaded ve-
locity ranges. The noise residual was 3.75 mK per channel in the
latter case, and 4.00 mK in the former (this fit was possibly limited
by the maximum 14th-order fit available in the SPLAT software).
Subtracting the narrow-fit baseline results in a detection with 6.1
sigma confidence, and integrated line signal of 106.4 ± 17.5 mK
km/s. The wide-fit baseline yields a 10.5 sigma detection of 200.0
± 19.1 mK km/s. Both velocity-intervals were 14 channels wide.
In the latter case, the remainder of the baselined spectrum could be
used to check for false-positives, and the maximum integrated sig-
nal found in a sliding 14-channel-wide window was 17.1 mK km/s.
The intensities are in a T∗
A antenna brightness temperature scale2.
The preferred spectrum results from the wide-fit baseline sub-
traction, because the depth of the central minimum is more consis-
tent with gas molecules on Keplerian orbits. Because of velocity
projection effects, the central minimum should always have a pos-
itive signal, even for an infinitesimally-thin ring. The toy models
shown adopt such a ring at 44 AU radius for the narrow-fit case,
compared to a belt at 35-45 AU in the wide-fit case. The former
model over-predicts the data by 2.4 sigma in the worst-case chan-
nel, and the latter by only 1.5 sigma. Also, the former case has a
residual (summed variance between model and data) of 207 mK2
compared to 164 mK2 expected from random noise, while the latter
fits within the noise (variance of 137 mK2 compared to random ex-
pectation of 192 mK2). The models are only intended to be indica-
tive, but are informative as to possible locations of the gas orbits.
For our fits to velocities in an edge-on disc around a 2.1 M(cid:12) host
star (Boccaletti et al. 2012), the radii thus point to molecules lying
inwards of the dust – for example, Donaldson et al. (2013) have
fitted a dust ring with a radius of around 110 AU.
There is a possible asymmetry between the red and blue sides
of the line profile, which in the preferred fit have integrated inten-
√
sities of 76.3 and 123.7 mK km/s respectively. The error on the
2 × 13.4 mK km/s, so this has only 2.5 sigma confi-
difference is
2 The Jy/K conversion factor for the JCMT is 15.6/ηa, with aperture ef-
ficiency ηa having standard values of 0.52 and 0.61 at 345 and 230 GHz
respectively.
Figure 3. CO J=2-1 spectrum for HD 32297. Top: results from fitting the
passband with wide and narrow (green-shaded) velocity ranges. Middle:
spectrum after subtracting the 3rd-order polynomial shown in red in the top
panel. Bottom: spectrum after subtracting the 14th-order polynomial shown
in brown in the top panel. The two models shown by the red curves in the
middle and bottom panels are for molecules on Keplerian orbits (see text).
At the half-power contour at 450 µm, the major:minor axis
ratio is 1.52 from this fit. The long diameter is 17.7 arcsec and the
minor axis is unresolved. This is consistent with the disc being seen
close to edge-on – Lieman-Sifry & Hughes (2015) find an inclina-
tion of 79.6±0.4◦ from their ALMA 850 µm image. They also iden-
tify dust out to 286 ± 7 AU from the star, while our 450 µm image
(after deconvolving the beam size in quadrature) gives a slightly
larger outer radius, approximately 6.7 arcsec or 395 AU at 59 pc.
Figure 2 shows our 850 µm image of HD 21997. The struc-
ture appear asymmetric, with most of the flux centred in a peak
slightly north of the star and only a 3-sigma counterpart to the
south. In contrast, Mo´or et al. (2013) found a compact symmetric
disc in their ALMA 886 µm image. However, they noted some ex-
tension along an axis at PA ∼ 20−25◦ in Herschel images at 70-100
µm. This agrees with the slightly east-of-north orientation seen in
the SCUBA-2 image. To reconcile the submillimetre datasets, there
could be significant flux beyond the half-power point of the ALMA
primary beam (>9 arcsec, or 650 AU). The SCUBA-2 integrated
flux in fact substantially exceeds the 2.7 ± 0.3 mJy in a 6-arcsec
c(cid:13) 2016 RAS, MNRAS 000, 1–8
4
J. S. Greaves et al.
Table 1. Literature data for the 27 debris-hosting A-stars within 125 pc that have been searched for CO, sub-divided into two age groups. The CII and OI lines
were searched for with Herschel; the Herschel Science Archive was used to check a few unpublished spectra. Blank entries indicate no observation has been
made, while u.l. denotes an upper limit. Stellar ages are from (1) Manoj et al. (2006); (2) TWA association: Ducourant et al. (2014); (3) LCC association:
Song et al. (2012); (4) Rhee et al. (2007); (5) Beta Pic Moving Group: Mamajek & Bell (2014); (6) Kalas (2005); (7) Argus association: Zuckerman & Song
(2012); (8) Tuc-Hor moving group: Bell et al. (2015); (9) Chen et al. (2014); (10) Mamajek (2012); (11) McDonald et al. (2012); (12) Monnier et al. (2012);
(M) Mo´or et al. (2015,2006).
star
alias
HD141569
HD109573
HD110058
HD131835
HD95086
HD121617
HD39060
HD172555
HD181296
HD32297
HD38206
HD85672
HD9672
HD3003
HD21997
HD102647
HD110411
HD17848
HD183324
HD182919
HD161868
HD95418
HD10939
HD216956
HD6028
HD158352
HD172167
HR4796
betaPic
etaTel
49Cet
beta3Tuc
HR1082
betaLeo
nuHor
gammaOph
Fomalhaut
HR6507
Vega
spectral
type
A0Ve
A0
A0V
A2IV
A8III
A1V
A6V
A7V
A0V
A0V
A0V
A0
A1V
A0V
A3IV/V
A3Va
A0V
A2V
A0V
A0V
A1V..
A1V
A1V
A4V
A3V
A7V
A0Va
distance
(pc)
99
73
107
123
90
120
19
29
48
112
75
93
59
46
72
11
36
51
61
73
32
24
62
7.7
91
60
7.7
age
(Myr)
7 (1)
8 (2)
10 (4)
16 (M)
17 (M)
17 (M)
23 (5)
23 (5)
23 (5)
30 (6)
30 (M)
30 (4)
40 (7)
45 (8)
45 (8)
50 (4)
90 (M)
100 (4)
140 (M)
200 (M)
200 (4)
320 (9)
350 (M)
440 (10)
500 (11)
600 (4)
700 (12)
CO?
CII?
OI?
incidences in age group
Yes
u.l.
u.l.
Yes
u.l.
u.l.
Yes
u.l.
u.l.
Yes
u.l.
u.l.
Yes
u.l.
Yes
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
Yes
Yes
u.l.
u.l.
Yes
u.l.
Yes
Yes
Yes
u.l.
u.l.
u.l.
u.l.
u.l.
Yes
Yes
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
u.l.
CO and/or C+: 7/16 (44%); OI: 3/10 (30%)
CO and/or C+: 0/11 (0%); OI: 0/3 (0%)
dence. (In the narrow-fit case, the respective quantities are 33.4 and
73.0 mK km/s, differing at the 2.25 sigma level.)
A scale, for nominal telescope efficiencies.
A CO J=3-2 line towards HD 32297 is tentatively seen in a
short JCMT observation with the HARP camera, from the archive.
The integrated signal is 226 ± 92 mK km/s (2.5 sigma) over a vstar±
9 km/s velocity range. However, Mo´or et al. (2011) reported a CO
3-2 upper limit from APEX which would translate to <∼ 150 mK
km/s on the JCMT T∗
The mean intensities across the JCMT 2-1 and 3-2 lines are
11.3 ± 1.1 mK (wide-fit baseline case) and 12.7 ± 5.2 mK, with
the latter value probably <∼ 8.4 mK for consistency with the APEX
data. This gives a 3-2/2-1 line ratio of ∼0.50 or less (after correction
for the different beam-filling factors for a presumed point-like disc).
In this case, the gas excitation temperature would be very low, at
below ∼5 K, but this only applies in a regime where both lines are
taken to be optically thin. This is not true in some other systems, as
noted below; see also Matr`a et al. (2015) for further discussion of
excitation of CO around an A-star.
3.3 CO census
From a master list of debris systems searched for CO rotational
transitions, we find 27 A-star hosts within 125 pc of the Sun (Table
1). The most recent systematic survey is that of Mo´or et al. (2015),
who also summarise previous work. The combined samples are
close to complete for fractional dust luminosities Ldust/L∗ > 10−4,
and have particularly targeted objects up to ages of 50 Myr. Our
new target HD 32297 is the sixth system where a CO line has been
detected. The earlier archetypes were found by Zuckerman et al.
(1995), using the IRAM 30 m telescope to discover CO 2-1 emis-
sion around HD 145169 and 49 Ceti. Corresponding CO 3-2 lines
were found with the JCMT (Dent et al. 1995, 2005). Mo´or et al.
(2011, 2015) have subsequently discovered CO transitions in the
HD 21997 and HD 131835 systems using the APEX 12m telescope.
Dent et al. (2014) have been the first to map a distribution of CO,
observing the 3-2 transition in the β Pic disc.
3.4 Properties of discs with gas
We now examine the seven C-hosting A-star systems, which have
ages ranging from around 7 to 45 Myr (Table 1). This group is
unified by the presence of CO and/or C+ – and often both signa-
tures, in support of the hypothesis that C+ may be a by-product of
CO photo-dissociation. (Observational limits may prevent this be-
ing universally seen: HD 21997 and HD 131835 have CO but only
upper limits on CII, while η Tel shows CII but not CO.) In contrast,
c(cid:13) 2016 RAS, MNRAS 000, 1–8
Gas and dust around A-stars at tens of Myr
5
Figure 4. Cumulative distributions for debris discs with carbon gas (solid
line) and without (dashed line), as functions of the stellar age in Myr.
OI persists for only about half as long, with the latest detection
being in the 23 Myr Beta Pic Moving Group (BPMG). Strong OI
emission is a characteristic of proto-planetary discs that are less
evolved than debris discs (Dent et al. 2013).
Figure 4 shows cumulative distribution functions for the sys-
tems with and without carbon gas, as a function of the stellar age.
At around 30 Myr, there is a divergence in the two populations,
with carbon-bearing systems ceasing by 50 Myr, but two-thirds of
the gas-less systems having greater ages. There do not seem to be
strong selection effects, as the list of systems from Table 1 is spread
quite evenly in log-age, reflecting a rather typical distribution for
detectable solid debris. The age distribution is not intrinsically that
of A-type stars, because debris fades with time (at a range of evolu-
tionary rates). Further, CO searches have concentrated on systems
with Ldust/L∗ > 10−4, which could introduce skews related to both
evolution and detectability by distance – however, any biases here
seem small. While less can be said about the older stars, the sam-
pling of sub-50-Myr systems is very complete within our distance
bound. Also, our median distance for stars ≤50 Myr old is 75 pc,
similar to 60 pc for older stars and 70 pc for systems with CO/C+
detections. We conclude that carbon-bearing gas appears in nearly
half of the well-studied group of debris systems around A-stars that
are up to 50 Myr old – well beyond the sub-10 Myr epoch associ-
ated with gaseous proto-planetary discs.
The CO-hosting discs in Table 1 have dust luminosities char-
acteristic of bright debris systems: Ldust/Lstar values range from
5 10−4 to 7 10−3 (Table 2), within the 'luminous' classification
of Mo´or et al. (2006). In contrast, proto-planetary discs are typ-
ified by Ldust/Lstar > 0.01. This empirical boundary has recently
been refined to criteria of R12 < 3 and R70 < 2000 (Wyatt et al.
2015), where R is the ratio of dust-to-photospheric emission and
the wavelength-subscripts are in microns. By this definition, HD
141569 may be an intermediate case, with R12 of 6 (Ldust/Lstar of
0.007), but still fairly distinct from gas-rich Herbig Ae stars. Hales
et al. (2014) present a discussion of recent CO detections in HAe-
type discs.
c(cid:13) 2016 RAS, MNRAS 000, 1–8
Figure 5. Correlations of gas and dust line fluxes (Table 2, but corrected
to a common distance of 100 pc). Triangles represent upper limits; η Tel
is omitted from the CO-correlation as it has limits on both axes (x ≤3.3, y
≤0.02).
4 DISCUSSION
4.1 Correlations
Figure 5 plots the CO and CII fluxes versus the 850 micron dust
emission. There are only 6-7 data points to examine for any trends,
but it appears that both gases rise in brightness more steeply
than the dust flux. For CO, the data can be fitted as F(CO) ∝
F(860µm)1.7. This is suggestive of a trend where the mass in CO
may scale with the production rate (i.e as the square of the dust
mass), but a more detailed treatment of gas excitation and opacity
is needed (see Mo´or et al. (2015), Figure 8, for the results of such
analysis).
Here we have explored trends in CO flux versus disc mass,
using the ProDiMo code (Woitke et al. 2016) to solve self-
consistently for chemistry and radiative transfer in steady-state.
The disc is assumed to have evolved into a 'late' proto-planetary
stage, with a gas-to-dust mass ratio of 10, and dust that has set-
tled downwards to a thin plane. While the disc chemistry produces
CO emission that rises more steeply than disc mass (Figure 6), the
absolute values of the line flux are too small compared to those
observed. In particular, dust masses (Mo´or et al. 2015) of ≤1 M⊕
are at or below the left end of the x-axis, but CO J=3-2 line fluxes
(Figure 5) lie at the top of the y-axis. This supports the hypothesis
that these are not simply evolved proto-planetary discs, but that the
CO molecules have a different origin. If their source is volatiles re-
leased from ices in collisions, then the starting point is not in fact
an H-rich gas disc, and also the gas-phase chemistry will probably
not be in steady state.
4.2 Asymmetries
Models predict that CO can have a very asymmetric spatial dis-
tribution, if the molecules originate from specific locations in the
disc. This could occur in any model in which CO is produced by
collisions between CO-ice bearing bodies and there are inhomo-
geneities in the density (and so collision rate) of the parent bodies.
Examples of such models include points resonant with a planet, or
101001.0.8.6.4.20KS–Test Comparison Cumulative Fraction PlotXCumulative Fraction6
J. S. Greaves et al.
Table 2. Continuum and line fluxes for the A-stars with CO detections, plus the CII-detected system η Tel. The 850-870 micron fluxes are from observations
here and in the literature, except for HD 32297 (estimated by interpolation over 500–1200 micron: Donaldson et al. 2013). The Herschel-detected lines of
CII and OI are at wavelengths of 158 and 63 microns respectively. All line fluxes are in units of 10−18 W/m2; continuum fluxes are in mJy (10−29 W/m2/Hz).
Uncertainties in brackets are 1-sigma and upper limits are 3-sigma.
HD 141569
HD 131835
beta Pic
η Tel
HD 32297
49 Cet
HD 21997
Ldust/L∗
F(850/870)
F(CO 3-2)
F(CO 2-1)
F(CII)
F(OI)
7e-3
12.6(±4.6)
0.14(±0.1)
0.03(-)
11.4(±1.8)
245(±5)
3e-3
8.5(±4.4)
0.032(±0.006)
≤0.018
≤0.5
≤1.5
3e-3
104(±10)
0.076(±0.008)
≤ 0.03
24.3(±0.4)
≈ 7
2e-4
≤ 14.4
≤ 0.083
1.7(±0.4)
3e-3
9(±3)
0.065(±0.027)
0.037(±0.004)
2.68(±0.72)
≤ 7.3
1e-3
14.8(±3.1)
0.069(±0.014)
0.013(-)
0.80(±0.17)
≤ 11
5e-4
6.3(±1.6)
0.038(±0.006)
0.018(±0.004)
<∼ 1.5
<∼ 2.7
Figure 6. ProDiMo code results. The host star is at 100 pc with Te f f = 8600
K, L∗ = 14 L(cid:12), X-ray flux of 1029 erg and UV flux of 0.1 L∗. The power-
law distribution of dust sizes has a -3.5 index over 0.01 to 300 mm. The gas
mass is ten times that of dust, and abundances are solar. The debris spans
30-100 AU in radius, with vertical height H(r) ∝ r1, to 10 AU height at 100
AU radius. The dust is settled to a thin mid-plane layer. The column density
through the disc scales as Σ(r) ∝ r−1, with exponential tapering inward of
40 AU and outward of 50 AU.
the site of the break-up of a massive planetesimal. The asymmetry
in the debris is enhanced for CO because of its short photodissocia-
tion timescale, so that the molecules are most concentrated at their
point of origin. Jackson et al. (2014) have modelled this scenario
for the case of a giant impact/break-up event, and Figure 7 illus-
trates how the observational outcome can vary for different view-
ing directions. The spectra are strongly asymmetric, but the shape
of the spectrum varies considerably with the (generally unknown)
viewing angle, complicating the interpretation. An additional com-
plication is that the simple model in Figure 7 only includes the Ke-
plerian orbital motion of the CO gas, and does not include the effect
of gas pressure, which may be significant at higher CO densities.
This would change the shape of the lines in Figure 7, but would not
remove the large scale asymmetry in the line profiles.
In Figure 8, whole-disc CO line-profiles are shown for four
Figure 7. Example of CO production for a giant impact model. The top-left
panel shows the spatial distribution of CO in a face-one view of the disk.
The bottom and top-right panels show spectra for the CO as viewed edge-on
from the bottom and right of the image respectively. The impact of an anti-
clockwise-orbiting progenitor occurs at 50 AU from a 1.5 solar-mass star,
producing a velocity dispersion of 0.3 times the circular Keplerian speed
(see Jackson et al. 2014). The CO decays exponentially with a timescale of
120 years after production.
systems. For β Pic, archival ALMA observations in CO J=2-1 were
used to generate disc-integrated spectra (with sensitivity up to ar-
cmin scales). For 49 Ceti and HD 141569, we used CO 3-2 lines
extracted from single-beam JCMT data3. The plots show tests for
asymmetry, where the line velocities around the central dip have
been swapped in sign, and used to generate a blue-red difference
spectrum. Where this residual is non-zero, the emission from the
the approach and receding sides of the disc is not the same, as for
example in many sightlines across the model disc in Figure 7.
3 For HD 141569, data from 2009 were used, with earlier archival spectra
from 1995 and 2001 showing similar profiles.
c(cid:13) 2016 RAS, MNRAS 000, 1–8
10-610-510-4Mdisk [Msun]10-610-510-410-310-210-1Fline (CO 3-2) [10-18 W/m2]Gas and dust around A-stars at tens of Myr
7
Table 3. Asymmetry measures for gas and dust. For CO, the ratio given is
of integrated intensity over two half-lines, with the mid-point defined as the
faintest line-centre channel. For dust images, the ratio given is of integrated
flux on either side of the star along the major axis; for the single dish data
these estimates only compare two beams. Data include: HD 141569 – CO
3-2 (JCMT), 0.87 mm dust (LABOCA); HD131835 – CO 3-2 (APEX),
0.87 mm (LABOCA); β Pic – CO 2-1 and 0.87 mm dust (ALMA); HD
21997 – CO 2-1, 3-2 (APEX), 0.85 mm dust (JCMT); HD 32297 – CO
2-1 (JCMT), 1.3 mm dust (CARMA); 49 Cet – CO 3-2 (JCMT, ALMA),
0.87 dust (ALMA). LABOCA images are from Nilsson et al. (2010); other
references are given above.
star
HD 141569
HD 131835
β Pic
HD 32297
49 Cet
HD 21997
CO line
asymmetry
1.0±0.2
∼1.3
≈2
1.6±0.3
∼1.8
≈1.0
mm dust
asymmetry
notes
>∼ 1.5
∼1.3
≈1.2
≈2
≈1.1
≈2.3
opt. thick CO
opt. thick CO?
CO more asym.
similar asym.
CO more asym?
opt. thick CO
SCUBA-2 (ALMA sees no low-velocity CO here). The line asym-
metry for HD 32297 is only at 2.5-sigma confidence, and the visual
appearance of asymmetry for HD 141569 is at the level of noise in
the difference spectrum.
Owing to the short dissociation time for CO molecules, whole-
disc spectra might be expected to be more skewed than the dust
distributions. Table 3 attempts to quantify this using the integrated
intensities of the blue/red sides of the CO lines, and the total dust
fluxes observed to either side of the star. The available data are
rather eclectic, but two systems do hint at the expected skewness
behaviour. Both β Pic and 49 Cet appear more asymmetric in CO
than in dust, using these measures, although the CO profile for the
latter is uncertain, as discussed above. HD 32297 and HD 131835
have spectral data too noisy to discriminate, while HD 21997 and
HD 141569 appear more asymmetric in dust. However, these two
systems appear to have optically thick CO lines such that saturation
would maximise intensity in both line-halves. The less abundant
isotopologue 13CO has recently been imaged in the 2-1 transition
(HD 21997: Kosp´al et al. 2013; HD 141569: P´ericaud et al. 2016),
and greater blue/red asymmetry begins to appear in these less sat-
urated lines. Both of these studies found large masses of CO, how-
ever, that may be inconsistent with sources in a comet belt. The
implications for the origin of the gas (collisional versus remnant
proto-planetary discs) are discussed by Zuckerman & Song (2012)
and Mo´or et al. (2015).
5 CONCLUSIONS
With the discovery of CO molecules around HD 32297, there are
now six A-type main-sequence stars where dusty debris is known
to be accompanied by carbon monoxide gas. In the age bracket
spanning 5-50 Myr, nearly half of the A-stars with debris in fact
exhibit carbon-bearing gas. This suggests a prolonged active epoch
where giant collisions release a burst of gas from frozen volatiles.
The time period is similar to that taken to assemble the Earth from
colliding planetesimals (e.g. Jacobson & Walsh 2015), so further
study may yield clues to how volatiles are folded into rocky planets
and their atmospheres.
As the photo-dissociation time for CO around A-type stars is
short, the gas profiles will tend to be asymmetric if debris origi-
Figure 8. CO whole-disc line profiles, and blue-red difference spectra (see
text). From top to bottom: HD 141569 observed in CO 3-2; β Pic in CO
2-1; HD 32297 in CO 2-1 (this work); 49 Cet in CO 3-2. Spectra of HD
21997 and HD 131835 (from APEX: Mo´or et al. 2011, 2015) are symmetric
within the noise, and are omitted. Y-axis units (not shown) are native to each
dataset; actual (dashed) and difference spectra are on the same scale within
each plot. X-axis units are velocity with respect to the line-centre minimum-
signal point.
From this analysis, we find that CO asymmetry is most marked
in β Pic. The high signal-to-noise line is asymmetric in both height
and width of the blue and red sides, reflecting the clump that has
been spatially resolved (Dent et al. 2013). For 49 Ceti, the overall
line shape appears similar to β Pic, but the nature of any asymmetry
is hard to characterise. Hughes (2014) has presented ALMA obser-
vations of the 49 Ceti disc (sampling scales only up to ≈10 arcsec)
and the line profile shows some differences to that from the JCMT.
Extended emission could thus be affecting the single dish spectrum,
although it appears uncontaminated by the northern object seen by
c(cid:13) 2016 RAS, MNRAS 000, 1–8
8
J. S. Greaves et al.
nates from one spatial location, with molecules mainly on one side
of the star. This is supported for 1-2 systems where the gas asym-
metry appears to exceed that of the dust. Models can explain a high
incidence of CO detection when the molecules are short-lived if,
for example, colliding debris repeatedly passes through the origi-
nal impact point. However, at least two of the discs are optically
thick in CO emission, suggesting the gas may not be collisional,
but instead represent a very prolonged proto-planetary disc phase.
ACKNOWLEDGMENTS
Data were obtained under JCMT project IDs MJLSD01 and
M13BU16. JSG and PW thank the ERC for funding for project Dis-
cAnalysis, under the grant FP7-SPACE-2011 collaborative project
284405. JPM is supported by a UNSW Vice-Chancellor's postdoc-
toral fellowship. MCW and LM acknowledge the support of the
European Union through ERC grant 279973.
The James Clerk Maxwell Telescope is operated by the East
Asian Observatory on behalf of The National Astronomical Ob-
servatory of Japan, Academia Sinica Institute of Astronomy and
Astrophysics, the Korea Astronomy and Space Science Institute,
the National Astronomical Observatories of China and the Chi-
nese Academy of Sciences (Grant No. XDB09000000), with addi-
tional funding support from the Science and Technology Facilities
Council of the United Kingdom and participating universities in
the United Kingdom and Canada. ALMA is a partnership of ESO
(representing its member states), NSF (USA) and NINS (Japan), to-
gether with NRC (Canada), NSC and ASIAA (Taiwan), and KASI
(Republic of Korea), in cooperation with the Republic of Chile.
The Joint ALMA Observatory is operated by ESO, AUI/NRAO and
NAOJ.
References
Bell C.P.M., Mamajek E.E., Naylor M., 2015, MNRAS 454,
593
Boccaletti A. et al., 2012, A& 544, A85
Chen C.H. et al., 2014, ApJS 211, 25
Dent W.R.F. et al., 2014, Science 343, 1490
Dent W.R.F. et al., 2013, PASP 125, 477
Dent W.R.F., Greaves J.S., Coulson I M., 2005, MNRAS 359,
663
Dent W.R.F., Greaves J.S., Mannings V., Coulson I.M., Walther
D.M., 1995, MNRAS 277, L25
Donaldson J.K., Lebreton J., Roberge A., Augereau J.-C.,
Krivov A.V., 2013, ApJ 772, 17
Ducourant C. et al., 2014, A&A 563, id.A121
Hales A.S., 2014, AJ 148, 47
Holland W.S. et al., 2013, MNRAS 430, 251
Hughes A.M., 2014, conference presentation archived at
http://www.ast.cam.ac.uk/talks/archive/3445
Hughes A.M., Wilner D.J., Kamp I., Hogerheijde M.R., 2008,
ApJ 681 626
Jackson A.P.., Wyatt M.C., Bonsor A., Veras D., 2014, MNRAS
440, 3757
Jenness T., Berry D., Chapin E., Economou F., Gibb A., Scott
D., 2011, ASPC 442, 281
Kalas P., 2005, ApJ 635, L169
K´osp´al ´A. et al., 2013, ApJ 776, 77
Kral Q., Wyatt M. C., Carswell R. F., Pringle J. E., Matr`a L.,
Juh´asz ´A., 2016, MNRAS, submitted
Lieman-Sifry J., Hughes A.M., 2015, AAS #225, id.349.18
Mamajek E.E., Bell C.P.M., 2014, MNRAS 445, 2169
Mamajek E.E., 2012, ApJ 754, L20
Maness H.L., Fitzgerald M.P., Paladini R., Kalas P., Duchene
G., Graham J.R., 2008, ApJ 686, L25
Manoj P., Bhatt H.C., Maheswar G., Muneer S., 2006, ApJ 653,
657
Matr`a L., Pani´c O., Wyatt M.C., Dent W.R.F., 2015 MNRAS
447 3936
McDonald I., Zijlstra A.A., Boyer M.L., 2012, MNRAS 427,
343
Monnier J.D et al., 2012, ApJ 761, L3
Montgomery S.L., Welsh B.Y., PASP 124, 1042
Mo´or A. et al., 2015, MNRAS 447, 577
Mo´or A. et al., 2013, ApJ 777, L25
Mo´or A. et al., 2011, ApJ 740, L7
Mo´or A. et al., 2006, ApJ 644, 525
Nilsson R., 2010, A&A 518, 40
Pani´c O. et al., 2013, MNRAS 435, 1037
Pani´c O., van Dishoeck E.F., Hogerheijde M.R., Belloche A.,
Gusten R., Boland W., Baryshev A., 2010, A&A 519, 110
P´ericaud J., di Folco E., Dutrey A., Augereau J.-C., Pi´etu V.,
Guilloteau S., 2016, IAUS 314, 201
Redfield S, 2007, ApJ 656 L97
Rhee J.H., Song I., Zuckerman B., McElwain M., 2007, ApJ
650, 1556 Rieke G.H. et al., 2005, ApJ 620, 1010
Roberge A. et al., 2013, ApJ 771, 69
Song I., Zuckerman B., Bessell M.S., 2012, AJ 144, id.8
Torres C.A.O., Quast G.R., Da Silva L., De La Reza R., Melo
C.H.F., Sterzik M., 2006, A&A 460, 695
Woitke P. et al., 2016, A&A 586, 103
Wyatt M.C., Pani´c O., Kennedy G.M., Matr`a L., 2015, Ap&SS
357, 103
Zuckerman B., Song I., 2012, ApJ 758 77
Zuckerman B., Forveille T., Kastner J.H., 1995, Nature 373, 494
This paper has been typeset from a TEX/ LATEX file prepared by the
author.
c(cid:13) 2016 RAS, MNRAS 000, 1–8
|
1601.05428 | 3 | 1601 | 2016-04-18T16:38:43 | Radio Emission from Red-Giant Hot Jupiters | [
"astro-ph.EP"
] | When planet-hosting stars evolve off the main sequence and go through the red-giant branch, the stars become orders of magnitudes more luminous and, at the same time, lose mass at much higher rates than their main-sequence counterparts. Accordingly, if planetary companions exist around these stars at orbital distances of a few AU, they will be heated up to the level of canonical hot Jupiters and also be subjected to a dense stellar wind. Given that magnetized planets interacting with stellar winds emit radio waves, such "Red-Giant Hot Jupiters" (RGHJs) may also be candidate radio emitters. We estimate the spectral auroral radio intensity of RGHJs based on the empirical relation with the stellar wind as well as a proposed scaling for planetary magnetic fields. RGHJs might be intrinsically as bright as or brighter than canonical hot Jupiters and about 100 times brighter than equivalent objects around main-sequence stars. We examine the capabilities of low-frequency radio observatories to detect this emission and find that the signal from an RGHJ may be detectable at distances up to a few hundred parsecs with the Square Kilometer Array. | astro-ph.EP | astro-ph |
Draft version October 13, 2018
Preprint typeset using LATEX style emulateapj v. 01/23/15
RADIO EMISSION FROM RED-GIANT HOT JUPITERS
Yuka Fujii1,2 David S. Spiegel3,4,5 Tony Mroczkowski6,7 Jason Nordhaus8,9
Neil T. Zimmerman10 Aaron R. Parsons11 Mehrdad Mirbabayi5 Nikku Madhusudhan12
1Earth-Life Science Institute, Tokyo Institute of Technology, Tokyo, 152-8550, JAPAN
2NASA Goddard Institute for Space Studies, New York, NY 10025, USA
3Analytics & Algorithms, Stitch Fix, San Francisco, CA 94103, USA
4Research & Development, Sum Labs, New York, NY 10001, USA
5Astrophysics Department, Institute for Advanced Study, Princeton, NJ 08540, USA
6National Research Council Fellow
8Department of Science and Mathematics, National Technical Institute for the Deaf, Rochester Institute of Technology,
7Naval Research Laboratory, 4555 Overlook Ave SW, Washington, DC 20375, USA
9Center for Computational Relativity and Gravitation, Rochester Institute of Technology, Rochester, NY 14623, USA
Rochester, NY 14623, USA
10Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
11Astronomy Department, University of California Berkeley, Berkeley, CA, USA and
12Astronomy Department, University of Cambridge, UK
Draft version October 13, 2018
ABSTRACT
When planet-hosting stars evolve off the main sequence and go through the red-giant branch,
the stars become orders of magnitudes more luminous and, at the same time, lose mass
at much higher rates than their main-sequence counterparts. Accordingly, if planetary
companions exist around these stars at orbital distances of a few AU, they will be heated up
to the level of canonical hot Jupiters and also be subjected to a dense stellar wind. Given
that magnetized planets interacting with stellar winds emit radio waves, such "Red-Giant
Hot Jupiters" (RGHJs) may also be candidate radio emitters. We estimate the spectral
auroral radio intensity of RGHJs based on the empirical relation with the stellar wind as
well as a proposed scaling for planetary magnetic fields. RGHJs might be intrinsically
as bright as or brighter than canonical hot Jupiters and about 100 times brighter than
equivalent objects around main-sequence stars. We examine the capabilities of low-frequency
radio observatories to detect this emission and find that the signal from an RGHJ may be
detectable at distances up to a few hundred parsecs with the Square Kilometer Array.
Keywords: planets and satellites: Jupiter -- Sun: evolution -- planetary systems -- stars:
evolution -- stars: AGB and post-AGB -- radio continuum: planetary systems
1. INTRODUCTION
Planets with strong magnetic fields may gener-
ate radio and/or X-ray emission when interacting
with energetic charged particles. It is well known
that Jupiter emits radio waves from its auroral re-
gion due to the cyclotron-maser instability (e.g. Wu
& Lee 1979; Zarka 1998; Treumann 2006). Exoplan-
ets could also generate radio emission through sim-
ilar mechanisms, depending on their intrinsic mag-
netic fields and the properties of surrounding plas-
mas, e.g. stellar wind particles and particles from
Io-like moons. This process could provide an avenue
to discover planets that are otherwise extremely dif-
ficult to find -- those orbiting highly evolved stars.
Unlike traditional methods of discovering exoplan-
ets, in which the planetary signal is a tiny pertur-
[email protected]
bation on light from the star, low-frequency plane-
tary radio emission might be an arena where planets
are vastly brighter than their stars. This paper is
an exploration of the surprisingly diverse range of
physical processes that lead to this emission and the
prospects for detecting radio emission from planets
around giant stars with current and near-future low-
frequency radio observatories.
Observations of radio emission from solar system
planets imply an empirical relation: the auroral ra-
dio power is proportional to the input solar wind en-
ergy going into each planetary magnetosphere. This
is commonly referred to as the "radiometric Bode's
law" (Desch & Kaiser 1984; Zarka et al. 2001). Ex-
trapolating this scaling to exoplanets, radio emis-
sion from exoplanetary systems has been examined.
Farrell et al. (1999), Zarka et al. (2001), and Lazio
et al. (2004) estimated that a few of the known ex-
2
oplanetary systems may have radio flux of ∼1 mJy
level, due to the small orbital distance of the plan-
ets. Stevens (2005) gave improved estimates of the
stellar mass-loss rate based on X-ray flux and re-
evaluated the radio flux of known exoplanetary sys-
tems. Griessmeier et al. (2005) took account of the
high stellar activity in the early stage of the plan-
etary system and proposed that the young system
would be a good candidate in which to search for ra-
dio emission. Griessmeier et al. (2007a,b) discussed
the effects of the detailed properties of stellar wind
in the proximity of the stars. They considered not
only the kinetic energy of the stellar wind but also
the magnetic energy of the stellar wind and coronal
mass ejection. Note that in these papers, the scaling
relations for the planetary magnetic field differ from
paper to paper. In Reiners & Christensen (2010),
the authors adopted a new scaling relation of plan-
etary magnetic fields based on Christensen et al.
(2009). Jardine & Collier Cameron (2008) modeled
the reconnection between the magnetic field of a
close-in planet and that of the host star to obtain
estimates that are not based on the "radiometric
Bode's law." Although observational searches for
these radio signatures are underway, no clear detec-
tion has been claimed (Bastian et al. 2000; George
& Stevens 2007; Smith et al. 2009; Stroe et al. 2012;
Hallinan et al. 2013; Murphy et al. 2015), while
there are some promising initial results (Lecavelier
des Etangs et al. 2013; Sirothia et al. 2014).
When stars less than ∼8 M(cid:12) evolve off the main
sequence, they evolve through the red-giant branch
(RGB) and the asymptotic-giant branch (AGB)
phases where their radii and luminosities increase by
orders of magnitude. Jovian planets in orbit around
such stars can migrate inward or outward due to
the interplay between tidal torques and the mass-
loss process on the post-main sequence (Nordhaus
et al. 2010; Kunitomo et al. 2011; Mustill & Villaver
2012; Spiegel 2012; Nordhaus & Spiegel 2013). Dur-
ing this time, such planets can be transiently heated
to hot-Jupiter temperatures ((cid:38)1000 K) at distances
out to tens of AU, depending on the star's mass;
such planets are termed "Red-Giant Hot Jupiters"
(RGHJs), though the term can refer to planets or-
biting either RGB and AGB stars (Spiegel & Mad-
husudhan 2012). Such planets are also subject to
interactions with a massive (but slow) stellar wind,
as the mass-loss rate of evolved stars is significant,
ranging from ∼10−8 M(cid:12) yr−1 to ∼10−5 M(cid:12) yr−1
with the highest values for AGB stars (e.g., Reimers
1975; Schild 1989; Vassiliadis & Wood 1993; Schoier
& Olofsson 2001; van Loon et al. 2005). On the
assumption that the radio emission is correlated
with the stellar wind, planetary companions around
evolved stars could also generate bright radio emis-
sion.
Based on this speculation, Ignace et al. (2010)
examined radio emission from known substellar-
mass companions around cool evolved stars. They
found that the low ionization fraction of the stellar
wind of evolved stars suppresses their radio emission
and leads to weak radio emission. In addition, they
considered a scenario in which post-bow shock heat-
ing could produce ionized hydrogen atoms. This
indicates a configuration of radio-wave generation
very different from that of the solar system, and
the lower-limit estimate, at least, is well below the
detection limit.
In this paper, we consider a further plausible
scenario, where the accretion of the massive stel-
lar wind onto the planet would emit UV and X-ray
photons that ionize the stellar wind in the vicinity of
the planet. The ionized stellar wind particles would
then interact with the planetary magnetic field in
the same way as the solar system planets do. Thus,
by extrapolating the "radiometric Bode's law," we
provide more optimistic estimates compared to the
previous result. In addition, we introduce two ma-
jor advances beyond the work of Ignace et al. (2010)
that are related to the observability of radio emis-
sion from RGHJs. First, we estimate the plasma
frequency cut-off of the stellar wind, which turns
out to be one of the major obstacles to detecting
radio emission from RGHJs. Second, we discuss
the planetary parameter space to search for radio
emission, employing scaling laws for the planetary
magnetic field, noting that the survey of exoplanets
around highly evolved stars is not complete.
In Section 2, we present the framework to ob-
tain the frequency and the flux of planetary auroral
radio emission, and we describe our models for the
stellar wind and planetary magnetosphere. Section
3 presents the estimates of the spectral radio flux
of RGHJs and compares the predictions with what
might be expected from canonical hot Jupiters as
well as those from Jupiter-twins. Section 4 gives
the prospects for the signal detection with the cur-
rent/future instruments. Estimates for the known
late-type (M-type) evolved stars are also included.
Finally, Section 5 concludes the paper with a brief
summary.
2. MODEL
In this section, we describe our framework to
estimate radio emission from RGHJs. First, we in-
troduce our scheme to compute the frequency and
intensity of planetary radio emission in Sections 2.1
and 2.2, respectively. Then, two ingredients for the
framework -- the strength of the planetary mag-
netic field and the properties of the stellar wind --
will be presented in Sections 2.3 and 2.4, respec-
tively.
In Section 2.5, we consider the ionization
around the planets, which is a crucial factor to de-
termine the efficiency of the interaction between the
planetary magnetosphere and the stellar wind.
2.1. Frequency of Radio Emission
As ionized electrons flow along planetary mag-
netic field lines, auroral radio waves are emitted at
the local cyclotron frequency. The upper limit is
around the cyclotron frequency of the planetary sur-
face magnetic field, νcyc, max:
(cid:18) B
(cid:19)
10 G
(1)
νcyc, max =
eB
2πmec
≈ 28 MHz
where B is the strength of the magnetic field at the
planetary surface, e and me are the charge and mass
of electrons, respectively, and c is the speed of light.
Radio emission from an exoplanet is observable
from the ground (on Earth) only when its frequency
is greater than both the plasma frequency of Earth's
ionosphere ν⊕
plasma and the maximum plasma fre-
quency along the line of sight νlos
plasma:
νcyc, max > ν⊕
plasma and νcyc, max > νlos
The plasma frequency may be expressed as
plasma
νplasma =
(cid:115)
= 8979 Hz ×(cid:16) ne
nee2
πme
cm−3
(cid:17)1/2
.
(2)
(3)
(4)
In the Earth's ionosphere, the electron number
density is less than 106 cm−3, which implies that
ν⊕
plasma ∼< 10 MHz. Along the line of sight, the maxi-
mum plasma frequency νlos
plasma is typically governed
by the density of stellar wind particles around the
planet, which will be specified in Section 2.4 below.
Another factor that affects the radio emission
is the plasma frequency at the site of radio-wave
generation. This is because the cyclotron-maser in-
stability as a mechanism to generate intense radio
emission is efficient when the in situ plasma fre-
quency is small in comparison to the local cyclotron
frequency (Treumann 2006). Future work will be
required to estimate the local plasma density and
examine this condition. In this paper, we assume
that this condition is satisfied at least at some re-
gion along the poloidal magnetic field lines.
2.2. Flux of Radio Emission
The auroral radio spectral flux of exoplanets ob-
served at the Earth, Fν, can be expressed by:
Fν =
Pradio
Ωl2∆ν
(5)
where Pradio is the energy that is deposited as auro-
ral radio emission of the considered frequency range,
3
Ω is the solid angle of the emission, l is the distance
between the target and the Earth, and ∆ν is the
frequency bandwidth.
We estimate the radio emission of exoplanets,
Pradio, simply by scaling the Jovian auroral radio
emission, Pradio, J, with the input energy from stel-
lar wind, in the same manner as Griessmeier et al.
(2005, 2007a,b). The scaling is based on the empir-
ical/apparently good correlation between the radio
emission intensity of solar system planets and the
input kinetic energy, Pinp, k, or the magnetic energy
of the solar wind, Pinp, m, (the "radiometric Bode's
law"; Desch & Kaiser 1984; Zarka et al. 2001), i.e.,
Pradio ∝ Pinp
Pinp, k = mpnv3 · πr2
Pinp, m = (B2
mag,
(cid:63)⊥/8π) v · πr2
mag ,
(6)
(7)
(8)
where mp is the proton mass, n is the number den-
sity of the stellar wind, v is the relative velocity
of the stellar wind particles to the planet, B(cid:63)⊥ is
the interstellar magnetic field perpendicular to the
stellar wind flow, and rmag is the distance from the
center of the planet to the magnetic stand-off point,
described below. It is not clear from observations of
the solar system planets which of the above two re-
lations (Eqs. 7 and 8) most accurately captures the
"true" relationship between input ingredients and
output radio power (Zarka et al. 2001). In this pa-
per, we assume that the radio emission scales with
the input kinetic energy (i.e., that Pradio ∝ Pk, inp,
as per Eq. 7) and consider the possible effects of a
dense stellar wind.
Note that if Eq. 8 is actually the better predictor
of radio power, it is difficult to estimate the emission
from RGHJs at this point, because magnetic fields
of evolved stars are not well constrained for highly
evolved stars; most observations set only upper lim-
its (e.g., Konstantinova-Antova et al. 2010, 2013;
Petit et al. 2013; Tsvetkova et al. 2013; Auri`ere et al.
2015).
In the case of the M-type giant EK boo,
a surface magnetic field ∼0.1-10 G has been mea-
sured. In this particular case, given the large stellar
radius (R(cid:63) ∼ 210R(cid:12)), the magnetic moment may be
about 103× larger than that of the Sun and there-
fore may also increase the planetary radio emission.
However, due to our present ignorance about the
strength of magnetic fields of evolved stars, we leave
the magnetic model of radio emission for RGHJs for
future work.
In reality, the total power of Jovian auroral emis-
sion varies greatly over time. The average is of the
order of 1.3 × 1010 W. During highly active peri-
ods, it averages more like 8.2 × 1010 W, and the
emission can reach powers as high as 4.5 × 1011 W
during peak activity (Zarka et al. 2004). Here, we
employ Pradio, J = 2.1×1011 W as a canonical value,
4
following Griessmeier et al. (2005, 2007b).
The magnetic stand-off point is where stellar
wind ram pressure and planetary magnetic pressure
are in approximate balance:
(cid:18) rmag
(cid:19)−6
Rp
mpnv2 ∼ B2
8π
.
(9)
The outflow ram pressures from heated planets are
negligible compared to these pressures, as discussed
in Appendix A. Therefore,
rmag ∼ Rp (8πmpn)
−1/6 v−1/3B1/3
(10)
The radius obtained for Jupiter from this equal-
ity is about half of the actual magnetospheric ra-
dius (Griessmeier et al. 2005). To estimate rmag for
RGHJs, we scale this radius according to the pa-
rameter dependence of equation (10).
For exoplanets, we assume that the solid angle
of the emission (Ω) is the same as that of Jupiter.
In reality, the solid angles of auroral radio emission
from Jupiter, Saturn, and Earth are ∼1.6, ∼6.3,
and ∼3.5 steradians, respectively (Desch & Kaiser
1984), which are on the same order and will not
significantly affect our order-of-magnitude estimate
of radio emission. The bandwidth, ∆ν, is assumed
to be proportional to the representative frequency
of the emission, which is the cyclotron frequency,
following Griessmeier et al. (2007b).
2.3. Assumptions for Planetary Magnetic Field
In order to obtain the frequency and the inten-
sity of the radio emission, we need to compute the
strength of the magnetic field at the planetary sur-
face, B. We do so simply by scaling the Jovian
magnetosphere at the surface BJ ∼ 10 G according
to the planetary mass and age, based on the scaling
relation described below.
Several scaling relations for planetary magnetic
field strength have been proposed (e.g. Busse 1976;
Russell 1978; Stevenson 1979; Mizutani et al. 1992;
Sano 1993; Starchenko & Jones 2002; Christensen
& Aubert 2006; Christensen et al. 2009), which are
summarized and compared with numerical simula-
tions in Christensen (2010). We employ the scaling
law proposed in Christensen et al. (2009) and used
in Reiners & Christensen (2010) to explore the evo-
lution of planetary magnetic fields:
dynamo ∝ fohm ρ1/3
B2
dynamo (F qo)2/3 ,
(11)
where Bdynamo is the mean magnetic field in the
dynamo region, fohm is the ratio of ohmic dissipa-
tion to the total dissipation, ρdynamo is the mean
density in the dynamo region, F is an efficiency
factor of order unity, and qo is the convected flux
at the outer boundary of the dynamo region (see
Christensen et al. 2009 for a comprehensive descrip-
tion). This scaling law is based on the assumption
that the ohmic dissipation energy is a fraction of
the available convected energy and was found to
be in good agreement with both the numerical ex-
periments (over a wide parameter space) and with
known objects from the Earth to stars. Here, fohm
and F are assumed to be constant for the bod-
ies considered in this paper. Dipole magnetic field
strength at the planetary surface, denoted by B, is
then scaled by
B ∝ Bdynamo
.
(12)
(cid:18) rdynamo
(cid:19)3
Rp
where rdynamo is the radius of the outer boundary
of the dynamo region.
The scaling law for Bdynamo, Equation (11), is
reasonable only for rapidly rotating objects. Un-
like canonical hot Jupiters, RGHJs are not tidally
locked to their host stars, so they probably have
the rapid rotation needed for Eq. 11 to be a use-
ful ansatz (Spiegel & Madhusudhan 2012). In this
paper, we assume that RGHJs indeed are rapidly
rotating so that they generate planetary magnetic
fields through the same mechanism as Jupiter.
In order to evaluate ρdynamo and qo, we need
a model of the internal planetary structure. We
consider Jupiter-like gaseous planets and assume
that the planetary radius is constant at Rp = Rp,J,
as numerical calculations show that the radii of
gaseous planets over the range of 0.1Mp,J < Mp <
10Mp,J (with core mass less than 10%) converge to
0.8Rp,J < Rp < 1.2Rp,J within 1 Gyr (Fortney et al.
2007; Spiegel & Burrows 2012, 2013). For the den-
sity profile, we follow Griessmeier et al. (2007b) and
assume a polytropic gas sphere with index n = 1,
which results in:
(cid:104)
(cid:18) πMp
4R3
p
(cid:19) sin
(cid:16)
(cid:105)
(cid:17) .
π r
Rp
π r
Rp
ρ[r] =
(13)
We determine the radius of the outer boundary
of the dynamo region, rdynamo, by assuming that
the hydrogen becomes metallic when ρ(r) exceeds
the critical density ρcrit = 0.7 g/cm3 (Guillot 2006;
Griessmeier et al. 2007b). The density of the metallic
core, ρdynamo is obtained by averaging the density in
the core. In the case of Jupiter, rdynamo,J = 0.85RJ
and ρdynamo,J = 1.899 g/cm3.
The scaling of convected heat flux at the outer
boundary, qo,
is obtained by dividing the age-
dependent net planetary luminosity, Lp, by the sur-
face area of the core region, i.e., 4πr2
dynamo. The
time-dependent luminosity is taken from equation
(1) of Burrows et al. (2001) (see also Marley et al.
2007). Ignoring the relatively weak dependence on
average atmospheric Rosseland mean opacity leads
to:
Lp ∼ 6.3 × 1023 erg
t
(cid:18)
(cid:19)−1.3(cid:18) Mp
(cid:19)−1.3(cid:18) Mp
(cid:19)2.64
(cid:19)−2
(cid:19)2.64(cid:18) rdynamo
(14)
MJ
.
Therefore, we have
qo ∼ qo,J
t
4.5 Gyr
4.5 Gyr
(cid:18)
MJ
rdynamo, J
(15)
2.4. Assumptions for Stellar Wind
The other key ingredient for radio emission is
the stellar wind. The number density of particles in
the stellar wind, n, can be expressed as
5
lar wind particles falling onto the planet, depends
on the stellar wind velocity, the infall velocity (i.e.,
the acceleration due to planet's gravitational field),
and the planetary orbital velocity. All three of
these terms might be of the same order of mag-
nitude: as described above, vsw is ∼30 km/s; the
infall velocity from the planetary gravity is ∼10-
25 km/s depending on the planetary mass (rang-
ing from 1MJto10MJ); and the orbital velocity is
10 − 30 km s−1 depending on the orbital distance
(ranging from 1 − 5 AU). In the next section, we
simply consider
.
∼ 10−1
v
v(cid:12)
(19)
a fiducial value for the normalization.
n =
M(cid:63)
4πa2mpvsw
,
(16)
2.5. Ionization of Stellar Wind Particles Around
the Planet
M(cid:63) is the stellar mass-loss rate, a is the
where
orbital distance from the star, mp is the proton
mass, and vsw is the velocity of the stellar wind.
M(cid:12) ∼ 2 × 10−14M(cid:12)/yr and
For the solar wind,
vsw ∼ 400 km/s (e.g., Hundhausen 1997).
The mass-loss rate of red giants is typically
M(cid:63) ∼ 10−8 − 10−7M(cid:12)/yr (Reimers 1975), and
the rate can be as high as 10−5M(cid:12)/yr during the
AGB phase (Schild 1989; Vassiliadis & Wood 1993;
Schoier & Olofsson 2001; van Loon et al. 2005).
Therefore, we have
M(cid:63)
M(cid:12)
∼ 106 − 109 .
(17)
∼(cid:112)2GM(cid:63)/R(cid:48)
The stellar wind velocity, vsw, becomes smaller
as the star evolves. The wind velocity is typically
of the order of the escape velocity at a distance of
several times the stellar radius (Suzuki 2007), i.e.,
(cid:63) is several times R(cid:63). For
a star with radius R(cid:63) = 100 R(cid:12), this results in
vsw ∼ 30 km/s. Therefore, a solar-mass red giant
with radius R(cid:63) = 100R(cid:12) produces a stellar wind
that is slower by an order of magnitude than that
of the Sun, v(cid:12).
(cid:63), where R(cid:48)
Based on equation (17), the number density of
the stellar wind (equation 16) is normalized as fol-
lows:
(cid:17)−2
n = 1.8 × 106 cm−3 ×(cid:16) a
(cid:33)(cid:18) vsw
(cid:32)
5 AU
M(cid:63)
×
10−8M(cid:12)/yr
30 km/s
As discussed in Ignace et al. (2010), as stars
evolve, the ionization fraction of the stellar wind di-
minishes to the order of ∼10−3 (Drake et al. 1987).
Since only charged particles interact with a plane-
tary magnetic field, this suggests inefficient interac-
tion with the planetary magnetosphere and hence,
low input energy for radio emission. However, for a
highly evolved star, the velocity of the stellar wind
eventually becomes slower than the escape veloc-
ity of the planetary companion and hence, stellar
wind particles will accrete onto the planets. Spiegel
& Madhusudhan (2012) considered Bondi -- Hoyle ac-
cretion where the accretion radius is
Racc =
2GMp
v2
rel
∼ 4 RJ
(cid:18) Mp
MJ
30 km s−1
vrel
(cid:19)(cid:16)
(cid:32)
(cid:19)3(cid:18) M(cid:63)
(cid:19)−2
(cid:19)
(cid:18) Mp
M(cid:12)
.
MJ
(20)
.
(21)
(cid:17)−2
(cid:33)
.
(22)
(23)
The accretion luminosity Lacc and temperature Tacc
are
Lacc ∼ 1025 erg s−1
M(cid:63)
10−8M(cid:12)/yr
(cid:18) Mp
MJ
×
Tacc ∼ 2 × 105 K
(cid:19)−1
.
(18)
The accretion onto planets, therefore, leads to emis-
sion of UV/X-ray photons whose characteristic en-
ergy ∼kBTacc exceeds the ionization energy of hy-
drogen, ERydberg = 13.6 eV (λ = 91.2 nm). There-
fore, UV/X-ray radiation from accretion will create
a local ionized region around the planet.
The velocity term in equation (7), which is the
relative velocity between the planet and the stel-
Let us consider the ionization profile around the
planet. We suppose a state where the ionization and
6
recombination rates are in equilibrium. Denoting
the ionization fraction by x, the equilibrium state
at a distance r from the planet may be represented
by
(24)
(25)
NX
4πr2 e−τ n(1 − x)σH[Ephoton] = (nx)2β[Te]
(cid:90) r
τ =
n(1 − x)σH[Ephoton]dr ,
RJ
where NX is the source rate of the photons that can
ionize hydrogen, Ephoton is the energy per photon,
and σH is the cross-section of H atoms for X-ray
photons. Per Verner et al. (1996), σH scales as
σH[Ephoton] ∼ 6.3 × 10−18 cm2 ·
(cid:18) Ephoton
(cid:19)−3
,
ERydberg
(26)
where β[Te] is the "class B" recombination coef-
ficient as a function of the electron temperature,
Te. We adopt the value at Te ∼ 104 K, β ∼
2.6 × 10−13 cm3/sec, in the following (Pequignot
et al. 1991); when the electron temperature is var-
ied from 103 to 105 K, β varies 1.5 × 10−12 − 3.2 ×
10−14 cm3 sec−1.
The source rate is obtained by counting the
number of photons with energy exceeding ERydberg,
which is approximately given by dividing the X-ray
accretion luminosity by the characteristic photon
energy produced:
where kB is Boltzmann's constant. As a result,
NX ∼ Lacc
kBTacc
(cid:26)4 × 1035 sec−1
4 × 1037 sec−1
NX ∼
,
(27)
(Mp = MJ)
(Mp = 10MJ)
.
(28)
For accretion onto very massive planets, the
characteristic photon energy is so high that the
released energetic electrons may also ionize other
atoms in the vicinity. The cross-section1 of hydro-
gen for electrons is ∼4×10−17 cm2 (Fite & Brack-
mann 1958), which implies a mean free path for ion-
ized electrons of ∼2RJ in the surrounding medium.
As a result, nearly 2/3 of released energetic elec-
trons will ionize a hydrogen atom within ∼2RJ, and
more than 99% will ionize a hydrogen atom within
∼10RJ.
In principle, photons with energy Ephoton have
the potential to ionize Ephoton/ERydberg hydrogen
atoms. To account for this, we consider two limiting
possibilities: After an ionizing collision, the energy
can (i ) be split evenly between the two electrons,
1 This is close to the geometric cross-section of the Bohr
radius.
or it can (ii ) go entirely into the kinetic energy of
one electron and not at all into that of the other (of
course, any split in between these extremes is pos-
sible, too). Note that conservation of momentum
and energy imply that the proton will not acquire a
significant fraction of the energy of the collision.2
Scenario (i ) implies a cascade where an elec-
tron with energy Ein ionizes an atom, producing
two electrons (an ionizing electron plus the released
electron) with energy (Ein − ERydberg)/2 for each.
For example, in an idealized case with Mp = 10MJ,
kBTacc ∼ 172 eV ∼ 12 ERydberg, a photoionization
could produce an electron with energy of (12− 1) =
11ERydberg, then a second ionization by that elec-
tron would produce two photons with energy of
(11 − 1)/2 = 5ERydberg, and the third ionization by
these two photons would produce four photons with
energy of (5 − 1)/2 = 2ERydberg, etc. The cascade
can proceed to the fourth order at the maximum.
Under scenario (ii ), the cascade proceeds with
the initial Ein electron leading to an electron with
kinetic energy Ein − ERydberg and another electron
with zero kinetic energy. Clearly, this cascade can
produce a maximum total of Ein/ERydberg free elec-
trons.
In reality, not all released electrons result in fur-
ther ionization interactions. If ξ represents the frac-
tion of released electrons that proceed to the next
ionization, then the number of ionized atoms Ni re-
leased through this cascade (i ) is
Ni = (1 − ξ) + 2ξ(1 − ξ) + 4ξ2(1 − ξ) + 8ξ3
= 1 + ξ + 2ξ2 + 4ξ3 .
(29)
Alternatively, cascade (ii ) leads to
1 − ξk
1 − ξ
Ni =
= 1 + ξ + ξ2 + ··· + ξk−1 ,
(30)
where k ≡ Ein/ERydberg is the maximum number of
ionizations for the given initial electron energy. This
limit leads to a value for Ni that is not dramatically
different from that of limit (ii ).
In Appendix B,
we estimate the Møller scattering cross-section and
show that cascade (ii ) -- unequal recoil energies --
is probably more realistic.
Ultimately,
NX in equation (24) is replaced by
NX → Ni NX .
(31)
For a 10MJ planet, Ni is probably approximately in
the range 5 -- 10.
2 Were it otherwise, we would have to take into account
what fraction of a rubber ball's kinetic energy is imparted to
the kinetic energy of the Earth when bouncing a ball.
which gives
rstromgren ∼
(cid:26)67 RJ
310 RJ
(Mp = MJ)
(Mp = 10MJ)
(35)
7
for a planet 5 AU from a red giant host. The
Stromgren radii for 1-MJ and 10-MJ planets are also
indicated in Figure 1.
Note that this is a very interesting and different
regime of the Stromgren sphere from the commonly-
considered case (photoionization around O/B-type
stars). Around RGHJs, the Stromgren sphere does
not delineate a sharp edge of ionization, because
of the smaller source rate and smaller photon-
hydrogen cross-section (in this case parameter "a"
in equation 13 of Stromgren 1939 is not small) on
account of X-ray photons interacting more weakly
with neutral hydrogen than UV photons near the
ionization limit. The Stromgren radii are indicated
by vertical arrows in Figure 1. Crucially, X-ray and
UV emission from accretion onto the planet should
ionize a significant fraction of the incoming stellar
wind, thereby allowing radio waves to be generated
from this interaction.
In reality, the temperature and luminosity based
on the Bondi -- Hoyle accretion (equations 22 and 23)
describes the situation only approximately. Pre-
cisely, only the neutral portion of the stellar wind
can accrete onto the planet without interacting with
the planetary magnetic field, and the ionized por-
tion would lose some energy at the bow shock be-
fore it accretes. On the other hand, the ionized
plasma interacting with the magnetic field could
have a cross-section larger than the Bondi -- Hoyle ac-
cretion radius. The detailed electromagnetic struc-
ture around RGHJs therefore requires elaborate nu-
merical simulations that are beyond the scope of
this paper. In the following sections, we aim to give
order-of-magnitude estimates of radio emission, ob-
servability, etc., which ought to be robust with re-
spect to uncertainties in the details of the ionization
process.
3. ESTIMATES
3.1. Planetary Magnetic Field and Frequency of
Radio Emission
Figure 2 shows the computed radius (rc) and av-
erage density (ρc) of the dynamo region as a func-
tion of planetary mass, as well as the heat flux
at the outer boundary of the core (qo), the esti-
mated strength of the planetary magnetic field (B),
and the corresponding cyclotron frequency (νcyc) as
functions of planetary mass and the age. Substitut-
ing equation (15) into equation (12), and given that
rdynamo does not change significantly, the magnetic
Figure 1. Profile of ionization fraction measured from the
surface of the planet, due to UV/X-ray from the accretion of
stellar wind onto the planet. Solid lines show the solutions
without the correction for the secondary ionization by ionized
electrons, and the dashed line shows the solution for a 10MJ
planet taking the correction into account with efficiency fac-
tor Ni = 6. The vertical arrows show the Stromgren radius
estimated simply using equation (35). The dotted vertical
lines indicate the location of the magnetic stand-off radii,
rmag, obtained using equation (40) below.
The ionization fraction x as a solution of equa-
tion (24) is shown in Figure 1. When τ ∼ nσH r
is much smaller than unity and thus the e−τ term
can be ignored (as is the case for Mp = 10MJ), the
solution is simply
−1 +(cid:112)1 + 4C[r]
x[r] =
2C[r]
C[r]≡ 4πnβ[Te]r2
N σH [Ephoton]
.
(32)
(33)
The solid lines show the ionization fraction corre-
sponding to no additional ionization by electrons
(i.e., ξ = 0, or Ni = 1) and the dashed line shows
the solution with Ni = 6 for a 10MJ planet. In the
figure, the vertical lines show the magnetic stand-off
radius obtained using equation (40) below. While
the photon rate is larger for more massive planets,
the strong dependence of the cross-section on pho-
ton energy (equation 26) leads to a decrease in the
radius of the ionized region. Nevertheless, a sub-
stantial amount of ionized plasma is expected around
the magnetic stand-off radius, despite the initially
low ionized fraction of the stellar wind.
The extent of the ionized region may also
be roughly estimated as the Stromgren radius
(Stromgren 1939):
(cid:33)1/3
(cid:32)
3
4π
NX
n2β
rstromgren =
,
(34)
100101102103distance from planet [RJ]0.00.20.40.60.81.0ionization fractionMp=MJMp=10MJrstromgrenrstromgrenrmagrmaga=5AU,10−8M⊙/yr8
Figure 2. Upper panels: radius and average density of the
dynamo region as a function of planetary mass. Lower pan-
els: evolutions of heat flux at the outer boundary, planetary
magnetic field, and maximum cyclotron frequency, for plan-
ets with varying masses (MJ: blue, 5MJ: green, and 10MJ:
red).
field is approximately:
(cid:18) Mp
(cid:19)1.04(cid:18)
(cid:19)−0.43
t
MJ
4.5 Gyr
B ∼ BJ
(36)
under this model, where BJ is the magnetic field
strength of Jupiter; in this paper, we roughly con-
sider BJ ∼ 10 G. Reasonably, the resultant val-
ues agree with Reiners & Christensen (2010), who
adopted the same scaling law for the planetary mag-
netic field; we show this figure just for completeness.
Note that the cyclotron frequency of Jovian planets
typically falls between 10 MHz and 1 GHz. In this
regime, there are a number of current and near-
future radio observatories including the Giant Me-
trewave Radio Telescope (GMRT), Low-Frequency
Array (LOFAR), Hydrogen Epoch of Reionization
Array (HERA), Square Kilometer Array (SKA),
and potential upgrades to the Very Large Array
(VLA) (see Section 4.3).
Since νcyc, max > ν⊕
plasma, the radio emission will
not be hindered by Earth's ionosphere cut-off. On
the other hand, it may experience opacity due to
the plasma of the stellar wind particles around the
planet. The maximum plasma frequency along the
line of sight, νlos
plasma, corresponds to that in the
vicinity of the planet, if the planet is on the near
side of its star to the Earth. Therefore, substitut-
ing equation (18) to ne in equation (4),
plasma ∼ 8979 Hz ×
νlos
× 1 cm3
(37)
(cid:33)1/2
(cid:19)−1/2
(cid:32)
= 12 MHz ×(cid:16) a
(cid:32)
×
M(cid:63)
10−8M(cid:12)/yr
4πa2mpvsw
M(cid:63)
(cid:17)−1(cid:18) vsw
(cid:33)1/2
5 AU
30 km/s
.
(38)
We can see emission only from where νcyc, max >
νlos
plasma. The detectable parameter space will be pre-
sented in more detail in the next section.
3.2. Flux of RGHJ Radio Emission in
Comparison with Canonical HJs
The magnetic stand-off radius (equation 10) may
be written as follows using the stellar mass-loss
rate:
rmag = rmag, J
(cid:18) B
(cid:19)1/3(cid:16)
(cid:33)−1/6(cid:18) v
(cid:32) M(cid:63)
BJ
a
(cid:17)1/3
(cid:19)−1/6
5.2AU
M(cid:12)
v(cid:12)
×
(39)
The typical value for RGHJs is found by substi-
tuting relevant values for stellar wind parameters
described in Section 2.4:
(cid:18) B
BJ
M(cid:63)
(cid:19)1/3(cid:16) a
(cid:17)1/3
(cid:33)−1/6(cid:18)
5AU
10−8M(cid:12)/yr
v
10−1v(cid:12)
rmag ∼ 14 RJ
(cid:32)
×
(cid:19)−1/6
(40)
where we employ rmag, J = 84 RJ (Joy et al. 2002).
Substituting equation (39) to equation (7), the
0.60.81.0rdynamo [RJ]024681012Mp [MJ]05101520ρdynamo [g/cm3]100102104106qo [erg/sec/cm2]101102103B [G]1081091010age [yr]100101102103νcyc,max [MHz]10MJ5MJMJ9
Figure 3. Radio flux in unit of Jy from a planetary companion to a red giant with mass loss rate 10−8M(cid:12)/yr (top) and
that to an AGB star with mass loss rate 10−5M(cid:12)/yr (bottom). The systems are located at 100 pc away. The doubly-hatched
regions show the parameter spaces where the planetary radio emission would not be observable at all because the maximum
frequency of the emission (cyclotron frequency at the planetary surface, νcyc) is below the plasma frequency cut-off, νlos
plasma.
The regions hatched with vertical lines show the parameter spaces where the frequencies of bulk radio emission is below νlos
plasma,
i.e., νlos
plasma > 0.1νcyc.
100101102orbital distance a[AU]RG (M=10−8M⊙/yr)1e-071e-061e-05radio flux in Jyfrom a target at 100pc, age 4.5Gyr100101102orbital distance a[AU]AGB (M=10−5M⊙/yr)1e-051e-041e-03100101102planetary mass Mp [MJ]101102103104νcyc,max[MHz]Pk, inp ≈ 140 Pk, inp, J
10
scaling of the radio emission is expanded as follows:
Pk, inp = mpnv3 · πr2
mag
(cid:18) B
(cid:19)2/3(cid:16)
(cid:33)2/3(cid:18) v
(cid:19)5/3
BJ
a
5.2 AU
(cid:17)−4/3
(41)
(42)
= Pk, inp, J
(cid:32) M(cid:63)
×
M(cid:12)
v(cid:12)
We may compare the radio emission power of
RGHJs at 5 AU with that of canonical hot Jupiters
set at 0.05 AU. In order to perform this comparison,
equation (42) can be re-normalized as follows:
≈ 14000 Pk, inp, J
(cid:32)
(cid:32)
×
×
10−8M(cid:12)/yr
10−1v(cid:12)
(43)
(for RGB stars' companions)
(cid:17)−4/3
(cid:19)5/3
(cid:17)−4/3
(cid:19)5/3
(cid:17)−4/3
v
BJ
M(cid:63)
5 AU
(cid:18) B
(cid:19)2/3(cid:16) a
(cid:33)2/3(cid:18)
(cid:19)2/3(cid:16) a
(cid:18) B
(cid:33)2/3(cid:18)
(cid:19)2/3(cid:16)
(cid:18) B
(cid:33)2/3(cid:18) v
(cid:19)5/3
M(cid:63)
BJ
BJ
a
v
0.05 AU
5 AU
≈ 490 Pk, inp, J
(cid:32) M(cid:63)
×
M(cid:12)
v(cid:12)
10−5M(cid:12)/yr
10−1v(cid:12)
(44)
(for AGB stars' companions)
≈ 2.4 × 10−5Jy
×
×
(45)
≈ 0.70 × 10−5Jy
×
×
×
×
×
×
BJ
BJ
v(cid:12)
M(cid:63)
M(cid:12)
5 AU
5 AU
100 pc
10−8M(cid:12)/yr
(cid:18) B
(cid:32)
(cid:18) B
(cid:19)−1/3(cid:16) a
(cid:17)−4/3
(cid:32) M(cid:63)
(cid:33)2/3(cid:18) v
(cid:19)5/3
(cid:18) d
(cid:19)−2
(cid:19)−1/3(cid:16) a
(cid:17)−4/3
(cid:33)2/3(cid:18)
(cid:18) d
(cid:19)−2
(cid:19)−1/3(cid:16) a
(cid:17)−4/3
(cid:33)2/3(cid:18)
(cid:19)−2
(cid:18) d
(cid:18) B
(cid:19)−1/3(cid:16)
(cid:17)−4/3
(cid:33)2/3(cid:18) v
(cid:32) M(cid:63)
(cid:18) B
(cid:32)
10−5M(cid:12)/yr
(cid:19)5/3
0.05 AU
100 pc
100 pc
5 AU
M(cid:63)
BJ
BJ
a
M(cid:12)
v(cid:12)
≈ 0.70 × 10−3Jy
(for a Jupiter-twin)
(cid:19)5/3
v
10−1v(cid:12)
(for RGB stars' companions)
(cid:19)5/3
(49)
v
10−1v(cid:12)
(for AGB stars' companions)
(47)
(48)
(50)
(for canonical hot Jupiters)
(for canonical hot Jupiters)
Here, we have normalized the magnetic field
strength of canonical hot Jupiters with BJ, consid-
ering the uncertainty of the magnetic fields of tidally
locked planets. Note that some models of plan-
etary magnetic field strength predict that tidally
locked planets have weaker magnetic fields due to
their slow rotation (e.g. Griessmeier et al. 2004). Al-
though the orbital velocity of canonical hot Jupiters
has been ignored in equation (45), the Keplerian
velocity at 0.05 AU around a solar-mass star is
∼130 km s−1, which results in only a <10% increase
of the relative velocity.
Using the expressions above, we find that the
radio spectral flux density observed at the Earth
is:
Fν =
Pradio
Ωd2νcyc
≈ 5.2 × 10−8Jy
(cid:18) d
(cid:19)−2
100 pc
(46)
Thus, RGHJs are expected to be intrinsically as
bright as the closest hot Jupiters. Compared with
the equivalent Jupiter-like planets around main se-
quence stars, the massive stellar wind of late red gi-
ants can increase the radio emission from planetary
companions by more than two orders of magnitude,
which allows us to explore 10 times more distant
systems, i.e., 1000 times more volume. This at least
partially compensates for the small population of
evolved stars.
Equation (48) gives a spectral flux density one
to two orders of magnitude smaller than the pre-
diction of equation (5) in Ignace et al. (2010) if we
assume the same set of fiducial values and full ion-
ization. This is primarily because their formula-
tion did not incorporate the effect of a compressed
planetary magnetosphere due to a massive stellar
wind, while the scaling law for the planetary mag-
netic field strength is also different.
Using this model for planetary magnetic field
strength, we may consider the radio spectral inten-
sity and maximum frequency of a given planetary
mass and age. Figure 3 shows contours of radio
spectral flux density from planetary companions at
4.5 Gyr with varying masses and orbital distances.
The target system is located at a distance of 100 pc
from the Earth; the flux is scaled by the distance as
a quadratic function.
The observable energy flux is limited by the
plasma cut-off frequency in the vicinity of the plan-
ets, given by equation (38). The doubly hatched
regions in Figure 3 indicate the regions of param-
eter space where νcyc, max < νlos
plasma, i.e., the ra-
dio emission from the planet cannot be observed
from the Earth. In reality, the peak of the auroral
radio emission does not always occur at the maxi-
mum frequency. Instead, it usually exists at lower
frequencies. Here, we also consider one more cri-
terion, νcyc, max > 10νlos
plasma as a conservative mea-
sure for the observability of the bulk of radio emis-
sion. In Figure 3, this conservative region is shown
as hatched regions with vertical lines. This indi-
cates that the plasma cut-off due to a dense stel-
lar wind is a significant obstacle to detecting plan-
etary companions around evolved stars. For a red-
giant system, the most promising targets are sys-
tems with massive companions with Mp (cid:38) 5MJ, al-
though systems with smaller companions at distant
orbits are also accessible. For an AGB-star system,
however, only very massive planets at distant orbits
are marginally detectable, because the stellar wind
is so dense that there is significant radio opacity due
to the plasma from the accreting stellar wind.
4. OBSERVABILITY
In this section, we discuss the observability of
the estimated radio emission. An obvious potential
obstacle is the intrinsic radio emission from host red
giant stars themselves; if they are bright relative to
the radio flux from the planets, it is significantly
more difficult to identify the planetary contribution.
We will see in Section 4.1 that radio flux from the
planets will probably be larger in a certain range
of parameter space. Then, in Section 4.2, we esti-
mate the radio flux from hypothetical Jupiter-like
planets around known M giants. In Section 4.3, we
examine the sensitivities and limitations of several
current and future radio instruments at relevant fre-
quencies. Finally, Section 4.4 notes the polarization
and the time variability of the radio emission as keys
to discern the signals from the background noise.
4.1. Intrinsic Radio Emission of Red-Giants Stars
Over the last four decades there have been nu-
merous radio continuum observations of nearby red
11
Figure 4. A cartoon of the radio emission spectra of an
RGHJ with 5MJ and the host red giants with 100 R(cid:12). The
spectrum of the RGHJ is modeled after Jovian radio spectra,
e.g. Figure 8 of Zarka et al. (2004) and Figure 2 of Griessmeier
et al. (2007a); the contribution from Io-DAM is not shown
here. The spectra of the host red giant are modeled simply
by extrapolating observed radio spectra above 1 GHz with a
power law. The shaded region represents the portion of the
spectrum below the plasma frequency of the stellar wind.
giants, many undertaken with the aim of under-
standing the extended atmospheres and mass-loss
mechanisms of K- and M-type giants (e.g., Newell
& Hjellming 1982; Knapp et al. 1995; Skinner et al.
1997; Lim et al. 1998; O'Gorman et al. 2013). Al-
most all observations have been carried out at fre-
quencies ≥5 GHz, probing thermal Bremsstrahlung
emission from the large, partially ionized envelopes
surrounding the giant stars (Drake & Linsky 1986).
The spectral index of this emission is of order unity,
with a range of reported values for various sources
between 0.8 and 1.6 (O'Gorman et al. 2013).
Only two published studies describe attempts to
detect continuum emission from single (non-binary)
red giant stars at frequencies below 1 GHz (L band);
both yielded null results. First, a program that ob-
served a sample of nine M-type giants at 430 MHz
with the Arecibo Telescope failed to detect any of
the sources down to flux density 10 mJy (Fix &
Spangler 1976). Later, with the Molonglo Observa-
tory Synthesis Telescope (MOST), a sample of eight
K- and M-type giants were observed at 843 MHz;
the upper limits of their flux densities were placed
at approximately 1 mJy (Beasley et al. 1992). Cur-
rent facilities could achieve far deeper sensitivity on
similar targets. For example, as listed in Table 1,
the GMRT at 150 MHz reaches a sensitivity better
than 0.5 mJy in under an hour. However, any re-
attempts to detect single red giants at wavelengths
near 1 meter remain to be presented in the litera-
ture.
The only red giants that have been detected at
meter wavelengths are those in interacting binary
systems like those of the RS CVn type. For ex-
12
ample, van den Oord & de Bruyn (1994) detected
plasma maser emission from II Pegasi at 360 MHz
and 609 MHz with the WSRT. Although single
dwarf-type main sequence stars are also known to
exhibit bright, coherent radio flares (Bastian 1990),
no analog to this non-thermal emission process has
been proposed to exist for evolved stars.
Altogether, we lack firm observational con-
straints on the brightness of M giants near 30-300
MHz frequencies.
However, assuming thermal Bremsstrahlung
emission continues to dominate in this wavelength
regime, we can extrapolate down from the reported
centimeter flux densities, using the specific spec-
tral index. For example, O'Gorman et al. (2013)
reported α Boo (R(cid:63) = 25.4R(cid:12), T = 4286 K,
d = 11.26 pc) at 1 GHz is about 70 µJy, and α Tau
(R(cid:63) = 44.2R(cid:12), T = 3910 K, d = 20.0 pc) at 3 GHz
is about 40 µJy. Using these data, we extrapolate
the radio flux from the host star F(cid:63) to the lower fre-
quency range with a simple power law as follows:
F(cid:63)(ν)∼ (0.4 − 1.4) × 10−5 Jy ×
(cid:19)−2(cid:18) R(cid:63)
(cid:19)2(cid:16)
(cid:18) d
(cid:17)α∗
(51)
,
ν
1 GHz
where the power index α∗ is of order unity.
100 pc
100 R(cid:12)
Figure 4 is a cartoon of the radio spectrum of a
planet with 5 MJ and that of the host red giant star
with 100 R(cid:12), both of which are placed at a distance
of 100 pc. Note that the thermal contribution from
the accretion is negligible in this figure. The spec-
tral shape is modeled by simply scaling the Jovian
radio spectrum (e.g. Figure 8 of Zarka et al. 2004).
The continuum lines from the star in the cases of
α∗ = 0.8 and 2.0 are shown as two cases. As Figure
4 indicates, with a reasonable range of spectral in-
dex, the radio emission from the star is smaller by
an order of magnitude than the expected radio flux
from RGHJs at frequencies below 300 MHz. For less
massive planets (with weaker magnetic fields), the
distinction is even clearer, because the peak flux is
larger and the peak frequencies are smaller.
Therefore, in the following, we ignore the radio
emission from the red giant as a noise source.
4.2. Radio Flux from Nearby M Giants
We now examine how realistic it is that RGHJ
radio emission might be detectable with current and
near-future instruments. In order to do so, we con-
sider if known red giant stars were to host planetary
companions and estimate the resulting radio flux
density. We use the list of M-type red giants from
Table 4 of Dumm & Schild (1998), which includes
estimates of stellar masses, radii, and effective tem-
peratures. Their data do not unambiguously dis-
tinguish RGB from early AGB stars, so probably
about 40% of the stars on the list are actually early
AGB stars. For each red giant, the mass-loss rate
is estimated using the improved Reimers' equation
(Reimers 1975) given by Schroder & Cuntz (2005,
2007):
(cid:32) L R
(cid:33)(cid:18) Teff
M
4000
(cid:19)3.5(cid:18)
1 +
g(cid:12)
4300g(cid:63)
(cid:19)
,
M(cid:63) ∼ 8×10−14 M(cid:12)/yr
(52)
where L = L(cid:63)/L(cid:12) is the stellar luminosity, R =
R(cid:63)/R(cid:12) is the stellar radius, M = M(cid:63)/M(cid:12) is the
stellar mass, and g(cid:63) is the surface gravity. The ve-
locity of the stellar wind is assumed to be the escape
velocity at ∼4R(cid:63):
(cid:114) 2GM(cid:63)
4R(cid:63)
v ∼
.
(53)
(cid:19)−2.5
(cid:18) M(cid:63)
(cid:19)1.075(cid:18) Mp
(cid:18) M(cid:63)
M(cid:12)
.
M(cid:12)
Mp, J
(54)
(cid:19)1.04
.
t ∼ 10 Gyr
The age of the system is taken to be our crude es-
timate of the main-sequence lifetime, as a function
of stellar mass. We simply assume:
Combining this relation with equations (1) and (36)
results in
νcyc, max ∼ 20 MHz
(55)
In addition, the distance to the system is obtained
based on the parallax data from Hipparcos data
sets3.
Using these data, we estimate the spectral
flux of radio emission by specifying planetary
mass and orbital distance of a hypothetical plan-
etary companion. We consider [A] a Jupiter-twin
whose maximum cyclotron frequency is νcyc, max ∼
20 (M(cid:63)/M(cid:12)) MHz, and [B] a larger Jovian planet
with Mp ∼ 10Mp whose maximum cyclotron fre-
quency is νcyc, max ∼ 200 (M(cid:63)/M(cid:12)) MHz.
Figure 5 displays the estimated radio flux from
planets [A] and [B] around M-type red giants within
300 pc. To span a reasonable range of orbital dis-
tances, we place the planets at 1 AU and at 5 AU
from their host stars. The shape of the marker
indicates whether the radio emission can escape
from the system (circles), i.e., νcyc, max > νlos
plasma,
or not (X's). Assuming Jupiter-mass planets at 1
AU (5 AU) which typically have the maximum cy-
clotron frequency at 10 -- 100 MHz depending on the
age, 148 (12) systems have potential to emit radio
waves at a flux density level above ∼10 µJy out
3
HIPPARCOS
http://www.rssd.esa.int/index.php?project=
to ∼100 -- 300 pc. Those systems with the highest
mass-loss rates, which are intrinsically the brightest
systems, would probably be unobservable because of
the plasma cut-off frequency of the stellar wind. For
more massive planets, with maximum cyclotron fre-
quency ∼100 -- 1000 MHz, the radio flux density is
lowered due to its dependence on B−1/3 (Equation
(47)), but is more likely to reach Earth because it is
less extinguished by plasma opacity. 10-MJ planets
at ∼1 AU may be detectable out to ∼200 pc.
A more realistic estimate of the expected num-
ber of detections may be evaluated by combining
these results with the probability for a star to have
a planet with a certain mass and orbit. Observa-
tions of main-sequence stars imply that the num-
ber of Jovian planets is in the range of several per-
cent to ∼10% of the number of stars (e.g. Cumming
et al. 2008; Cassan et al. 2012). Assuming the same
population for RG systems, the expected number
of actual detections from RGHJs would be of or-
der unity. A more detailed examination using the
empirical planet mass function by Cumming et al.
(2008) is given in Appendix D.
13
4.3. Detectability with Current and Near-future
Low-frequency Radio Instrumentation
We consider in this section several current and
near-future radio observatories that operate at fre-
quencies ν (cid:46) 350 MHz and can reach continuum
sensitivities less than 1 mJy in a few hours. The
results are summarized in Table 1.
The primary limitation to detectability is a tele-
scope's sensitivity, characterized by the root mean
square (RMS) noise fluctuations in the sky and re-
ceiver. The RMS noise σRMS in an interferometric
observation with integration time τ , bandwidth ∆ν,
effective area Aeff (≈ 0.7π(D/2)2 for interferome-
ters comprising antennas with dish size D), and N
antennas is (see e.g. Condon & Ransom 2016)
(cid:112)∆ν τ N (N − 1)
2kTsys
(cid:18) 1026Jy
(cid:19)
Wm−2Hz−1
.
σRMS =
Aeff
(56)
Here, Tsys is the blackbody-temperature equivalent
of the total system noise, which is the sum of the
instrument or receiver noise temperature Trx and the
noise contribution from the sky Tsky. We estimate
Tsky using the low-frequency sky noise temperature
fit from Rogers & Bowman (2008):
Tsky = T150
+ TCMB
(57)
(cid:16)
ν
150 MHz
(cid:17)−β
where T150 = 283.2 K is the Tsky at 150 MHz,
β = 2.47, and TCMB = 2.73 K is the contribution
from the cosmic microwave background (CMB).
While this relation was fit to data taken in the 100 --
200 MHz range, the authors found it to be consistent
with published measurements from 10 to 408 MHz.
An additional factor that affects detectability is
source confusion, the imaging noise associated with
unresolved interloping radio sources in an observa-
tion. Source confusion is characterized by the stan-
dard deviation σc in the surface brightness of an
image due to one or more unresolved sources in the
beam solid angle (for a review, see Condon 1974;
Condon et al. 2012). To estimate σc, reported in
Table 1, we use the relations provided by Condon
& Ransom (2016), reproduced below:
(cid:16)
(cid:16)
σc ≈ 0.2
σc ≈ 2.2
ν
1 GHz
ν
1 GHz
(cid:17)−0.7(cid:18) θ
(cid:17)−0.7(cid:18) θ
1(cid:48)
(cid:19)2
(cid:19)10/3
1(cid:48)
, (θ > 10(cid:48)(cid:48))
(58)
, (θ < 10(cid:48)(cid:48)). (59)
Figure 5. Radio flux from hypothetical RGHJs of 1MJ (up-
per panel) and 10MJ (lower panel) at 1 AU (blue) and 5 AU
(black). X symbols indicate that the radio emission is not
observable because cyclotron frequency is less than plasma
density around the planet.
As Table 1 illustrates, source confusion is likely
to exceed the 2-µJy detection limit of a typical
RGHJ source for most telescopes under considera-
tion. However, source confusion can be mitigated
through differencing observations that cancel out
background sources. Potential axes for such cancel-
lation are polarization and time. The source con-
50100150200250300distance [pc]10-610-510-410-3radio emission [Jy]for Mp=1MJ planet1AU5AU50100150200250300distance [pc]10-610-510-410-3radio emission [Jy]for Mp=10MJ planet1AU5AU14
Detectability of Hot Jupiters with Current and Future Radio Telescopes.
Table 1
Instrument
LOFAR-LBA
LOFAR-HBA
HERA
GMRT
SKA-Low (part)
SKA-Low (full)
VLA-LOBO
Band
(MHz)
30-80
110-200
50-250
130-190
50-200
50-350
50-350
a
σRMS
(µJy bm−1)
1100
180
76
162
5.9
1.8
77
τ b
(hr)
230000
6500
1200
5400
7.2
0.7
1200
Resolution
((cid:48)(cid:48))
10
4.1
410
16.5
7.6
4.8
8.6
c
σc
(µJy bm−1)
57
1.1
36000
57
9.7
1.5
10.4
a RMS noise in a 1-hour observation, computed for imaging of the full band
reported in this table (Equation 56).
b The integration time in hours required to reach an RMS of 2 µJy (i.e. 5-σ
detection of a typical source at 100 pc with 10 µJy bm−1 average flux).
c Source confusion RMS contribution.
fusion limit is dominated by active galactic nuclei,
which have low polarization fractions (cid:46) 2.5% (Stil
et al. 2014) and slowly vary with time. Therefore,
both the circularly polarized nature of the cyclotron
emission from planets and the time variability (dis-
cussed in §4.4) partially mitigate the source confu-
sion limit, allowing planetary radio emission, like
other time-variable sources, to be detected below
the imaging confusion limit. We therefore expect
confusion not to be a strong limitation to the de-
tectability of radio emission from RGHJs.
The LOFAR operates as two separate arrays de-
fined by their high and low bands, differing con-
figurations, and antenna designs, which are called
respectively the High Band Antennas (HBA) and
Low Band Antennas (LBA). LOFAR-LBA operates
in the 10 -- 90 MHz range, and LOFAR-HBA oper-
ates in the 100 -- 240 MHz range (van Haarlem et al.
2013). We consider it as two separate instruments
here. LOFAR-LBA's most sensitive band is cen-
tered at 60 MHz, and LOFAR-HBA's is at 150 MHz.
Given the strong frequency dependence of the effec-
tive aperture of a dipole antenna, and that the re-
ceiver noise temperature for LOFAR Trx ∼ Tsky, we
use the LOFAR Image Noise calculator4, which is
based on SKA Memo 1135, to obtain the sensitivity
and imaging parameters in Table 1.
The HERA6 is a telescope array under construc-
tion in South Africa that will ultimately consist of
352 14-m parabolic dishes, operating from 50 to
250 MHz (Pober et al. 2014). Unlike the other tele-
scope observatories considered here, HERA is a ded-
icated experiment for making power spectral mea-
surements of cosmic reionization over wide fields of
view. Similar to its progenitor, the Precision Array
to Probe the Epoch of Reionization (PAPER; Par-
sons et al. 2010), HERA's dishes are deployed in
4 Heald; http://www.astron.nl/~heald/test/sens.php
5 Nijboer, Pandey-Pommier, & de Bruyn; http://www.
skatelescope.org/uploaded/59513_113_Memo_Nijboer.pdf
6 http://reionization.org/
an extremely compact configuration (Parsons et al.
2012) and statically point toward zenith. While
these features make HERA less ideal for targeted
observations, HERA's substantial collecting area,
long observing campaigns, and wide field of view
make it a powerful survey instrument.
The GMRT is an array comprising 30 25-meter
antennas, operating at frequencies 130 -- 2000 MHz.
The lowest band (130 -- 190 MHz) exhibits receiver
noise Trx comparable to the sky noise. To be con-
servative, we use the GMRT online calculator.7
The low-frequency component of the SKA, re-
cently rebaselined to have 0.4 km2 collecting area in
130,000 elements with baselines extending to 65 km,
will operate from 50 to 350 MHz.8 We estimate the
noise in the lower half of the band (ν < 200 MHz),
appropriate for lower mass ((cid:46) 10MJ) separately
from the full band, which could be used to de-
tect ∼30-MJ objects. We again adopt the Rogers
& Bowman (2008) relation for Tsys (Equation (57))
and assume Trx ≈ 60 K. We caution that these
assumptions are optimistic, and the time estimates
could in reality be a factor of a few worse.
The low-frequency receivers (28 -- 80 MHz; 4-
Band) in the VLA have Trx ∼ 260 K, similar to
the receiver noise temperatures for the Long Wave-
length Array (LWA; Hicks et al. 2012). As noted in
Hicks et al. (2012), this noise level is subdominant
to Tsky by at least -6 dB. Future upgrades to the
VLA such as the LOw Band Observatory (LOBO;
Kassim et al. 2015) are being considered, and could
potentially cover the full 30 -- 470 MHz band. We
estimate the sensitivity of LOBO assuming the re-
ceiver noise extrapolates linearly from Trx ≈ 60 K
(internal VLA memo) in P-Band (230 -- 470 MHz)
to Trx ≈ 260 K in 4-band, and consider just the 50 --
http://gmrt.ncra.tifr.res.in/~astrosupp/obs_
setup/sensitivity.html
7
8
worlds-largest-radio-telescope-near-construction/
https://www.skatelescope.org/news/
350 MHz portion of the full 30 -- 470 MHz band.
The noise estimates shown in Table 1 indicate
that, at the frequencies relevant for the detection
of radio emission from Jupiter-size planets, current
instrument sensitivities are generally too low. On
the other hand, for more massive exoplanets (Mp ∼
10MJ) where the bulk of the radio emission would
exist above 50 MHz, the low-frequency component
of SKA will be able to detect the signal within half a
day from planets with Mp ∼ 10 MJ at ∼1 AU within
∼200 pc, or those at ∼5 AU within ∼100 pc. VLA-
LOBO could also work for a few nearby systems.
4.4. Discriminating the Signal from the
Background
Since the confusion limit is on the same order
as the signal (Table 1), it is crucial to consider the
ways to discriminate the signal from these sources.
There are two key features that makes planetary ra-
dio emission distinct: the circular polarization and
the time variability (Hess & Zarka 2011; Zarka et al.
2015).
Planetary auroral radio emission is nearly com-
pletely circularly polarized (e.g., Dessler 1983, and
references therein). In contrast, active galactic nu-
clei -- the major confusion sources -- have low po-
larization fractions (cid:46)2.5% (Stil et al. 2014).
The other key feature is the significant time vari-
ability of the planetary radio emission. The factors
that influence time variability are listed below:
Modulation due to planetary spin rotation ( ∼a few
hours) -- Radio bursts from Jupiter tend to be cor-
related with its spin phase (e.g. Dessler 1983) due
to the misalignment between the magnetic axis and
the spin axis. Similarly, high-temporal-resolution
observations of an RGHJ system could reveal peri-
odicity at the planet's spin frequency.
Modulation due to the presence of satellites (∼a few
days) -- Jupiter's radio bursts are affected by the
location of its moon Io, as well (Dessler 1983), be-
cause Io disturbs the surrounding electromagnetic
field. Likewise, if an RGHJ has a moon that emits
plasma, then the radio flux from such a system is
likely to be modulated by the orbital motion of the
moon (see, e.g., Noyola et al. 2014).
Variability of the stellar wind (∼a few months?) -- If
the stellar wind passing the planet is variable, any
modulation of the plasma density would also lead
to time variability of planetary radio emission. Such
variability might be expected to occur on timescales
of several months to half a year, as many M giants
have semi-regular periodicities on the order of a few
hundred days (Kiss et al. 1999).
15
Modulation due to orbital motion of the planet (∼a few
years) -- If the planetary orbit is close to the edge-
on configuration, the radio emission is likely to dis-
appear or significantly dim at certain orbital phases,
due to the increased plasma cut-off frequency along
the path and/or secondary eclipse (Lecavelier des
Etangs et al. 2013). The plasma cut-off frequency
increases at some phases because, when the planet
is on the far side of the star from Earth, the path
of the radio emission toward Earth comes through
the vicinity of the star, where the plasma density
may be high enough to be partially or completely
opaque to the emission, depending on the ionization
fraction of the stellar wind.
5. SUMMARY AND DISCUSSION
In this paper, we estimate the radio bright-
ness of distant "hot Jupiters" around evolved stars
(RGHJs). Unlike Ignace et al. (2010), we consider
that UV/X-ray photons from accretion onto the
planets partially ionize the stellar wind in the vicin-
ity of planets, which would otherwise be mostly neu-
tral for a cool star's wind. This process yields the
free electrons that are crucial to producing the ra-
dio emission. Based on such a picture, the dense
stellar wind of an RGB or AGB star would interact
with the planetary magnetic field and would add
power in the form of kinetic energy into the mag-
netosphere of an RGHJ. We find that the intrin-
sic brightness of radio emission from RGHJs could
be comparable to or greater than that of canonical
hot Jupiters in close-in orbits around main-sequence
stars and >100 times brighter than distant Jupiter-
twins around main-sequence stars. This implies
that they can be searched for 10 times further away
or in a volume 1000 times as large. This can com-
pensate for the rareness of the evolved stars at least
partly. Thus, RGHJs will serve as reasonable tar-
gets in future searches for exoplanetary radio emis-
sion.
A major obstacle to observing this radio emis-
sion is the plasma cut-off frequency of the (ionized)
stellar wind. Due to the great density of the stellar
wind, the cut-off frequency is as high as ∼12 MHz
for typical red giants and ∼400 MHz for typical
AGB stars, making planetary-mass companions to
AGB stars difficult or impossible to see via their
auroral radio emission. The most promising targets
are massive planetary companions (Mp (cid:38) 5MJ) to
red giant stars, with magnetic fields stronger than
that of Jupiter (and, therefore, higher cyclotron fre-
quencies). Similarly, if the plasma frequency in the
region where the bulk of the radio emission is gen-
erated is large relative to the cyclotron frequency,
this could pose a serious obstacle to observability.
The radio flux from a system with a Jovian ex-
oplanet at a distance of 100 pc would typically be
16
on the order of ∼10 µJy. Such signals would be
detectable with SKA within half a day. For a few
nearby systems, a possible upgrade to the VLA such
as LOBO would also provide reasonable integration
times and resolving power. In both cases, it is crit-
ical to consider polarization and/or time variability
of the planetary radio flux, in order to separate it
from confused sources in the same resolution ele-
ment.
The radio emission from RGHJs may soon prove
to be a valuable tool for surveying exoplanets in
a region of parameter space around highly evolved
stars where traditional exoplanet-discovery meth-
ods, such as transit and radial-velocity observations,
are less sensitive. Once this planetary radio emis-
sion is detected, it will provide a unique approach
for studying RGHJ properties. As the auroral ra-
dio flux is significantly modulated in association
with several physical processes described in Section
4.4, the time variability is useful to infer the plan-
etary spin rotation period or the presence of satel-
lites. If spectra are obtained, the upper cut-off fre-
quency can tell us the magnetic field strength at
the planetary surface, which, when combined with
constraints on the planetary mass via, e.g., radial-
velocity observations, tests the proposed scaling law
for the planetary dynamo. Additionally, the lower
cut-off frequency of the RGHJ emission could probe
the stellar wind properties at the planetary orbit.
And finally, the planetary radio power in such an ex-
treme environment as an RG's massive stellar wind
will provide valuable data points to test the empir-
ical "radiometric Bode's law."
Acknowledgments
Y.F. is supported from the Grant-in-Aid No.
25887024 by the Japan Society for the Promotion
of Science and from an appointment to the NASA
Postdoctoral Program at NASA Goddard Institute
for Space Studies, administered by Oak Ridge Af-
filiated Universities. D.S.S. gratefully acknowledges
support from a fellowship from the AMIAS group.
J.N. acknowledges support from NASA grant HST
AR-12146.04-A and NSF grant AST-1102738. We
thank David Hogg for encouraging us to pursue
calculations of exoplanetary radio emission. We
thank Greg Novak for helpful conversations. We
thank Jake VanderPlas for developing the XKCD-
style plotting package for matplotlib, and acknowl-
edge its use. A.R.P. is grateful for support from
NSF grants 1352519 and 1440343. Y.F. greatly ac-
knowledges insightful and helpful discussions with
Tomoki Kimura and Hiroki Harakawa. The portion
of this research for which T.M. is responsible was
performed under a National Research Council Re-
search Associateship Award at the Naval Research
Laboratory (NRL). Basic research in radio astron-
omy at NRL is supported by 6.1 Base funding.
REFERENCES
Auri`ere, M., Konstantinova-Antova, R., Charbonnel, C.,
et al. 2015, A&A, 574, A90
Bastian, T. S. 1990, SoPh, 130, 265
Bastian, T. S., Dulk, G. A., & Leblanc, Y. 2000, ApJ, 545,
Beasley, A. J., Stewart, R. T., & Carter, B. D. 1992,
1058
MNRAS, 254, 1
2001, RvMP, 73, 719
Burrows, A., Hubbard, W. B., Lunine, J. I., & Liebert, J.
Busse, F. H. 1976, PEPI, 12, 350
Cassan, A., Kubas, D., Beaulieu, J.-P., et al. 2012, Natur,
481, 167
Christensen, U. R. 2010, SSRv, 152, 565
Christensen, U. R., & Aubert, J. 2006, GeoJI, 166, 97
Christensen, U. R., Holzwarth, V., & Reiners, A. 2009,
Natur, 457, 167
Condon, J. J. 1974, ApJ, 188, 279
Condon, J. J., & Ransom, S. M. 2016, Essential Radio
Astronomy (Princeton Series in Modern Observational
Astronomy, Princeton University Press, 2016)
Condon, J. J., Cotton, W. D., Fomalont, E. B., et al. 2012,
Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008,
ApJ, 758, 23
PASP, 120, 531
Desch, M. D., & Kaiser, M. L. 1984, Natur, 310, 755
Dessler, A. J. 1983, Physics of the Jovian magnetosphere
(Cambridge: Cambridge Univ. Press)
Drake, S. A., & Linsky, J. L. 1986, AJ, 91, 602
Drake, S. A., Linsky, J. L., & Elitzur, M. 1987, AJ, 94, 1280
Dumm, T., & Schild, H. 1998, NewA, 3, 137
Farrell, W. M., Desch, M. D., & Zarka, P. 1999, JGR, 104,
14025
659, 1661
Fite, W. L., & Brackmann, R. T. 1958, PhRv, 112, 1141
Fix, J. D., & Spangler, S. R. 1976, ApJ, 209, 503
Fortney, F. J., Marley, M. S., & Barnes, J. W. 2007, ApJ,
George, S. J., & Stevens, I. R. 2007, MNRAS, 382, 455
Griessmeier, J.-M., Motschmann, U., Mann, G., & Rucker,
H. O. 2005, A&A, 437, 717
Griessmeier, J.-M., Preusse, S., Khodachenko, M., et al.
2007a, P&SS, 55, 618
Griessmeier, J.-M., Zarka, P., & Spreeuw, H. 2007b, A&A,
475, 359
A&A, 425, 753
Griessmeier, J.-M., Stadelmann, A., Penz, T., et al. 2004,
Guillot, T. 2006, in Saas-Fee Advanced Course 31:
Extrasolar planets, ed. D. Queloz, S. Udry, M. Mayor,
W. Benz, P. Cassen, T. Guillot, & A. Quirrenbach,
243 -- 368
Hallinan, G., Sirothia, S. K., Antonova, A., et al. 2013,
ApJ, 762, 34
Hess, S. L. G., & Zarka, P. 2011, A&A, 531, A29
Hicks, B. C., Paravastu-Dalal, N., Stewart, K. P., et al.
2012, PASP, 124, 1090
Hundhausen, A. J. 1997, in Cosmic Winds and the
Heliosphere, ed. J. R. Jokipii, C. P. Sonett, & M. S.
Giampapa (Tucson, AZ: Univ. Arizona Press), 259
Ignace, R., Giroux, M. L., & Luttermoser, D. G. 2010,
MNRAS, 402, 2609
Jardine, M., & Collier Cameron, A. 2008, A&A, 490, 843
Joy, S. P., Kivelson, M. G., Walker, R. J., et al. 2002,
Kassim, N. E., Clarke, T., Helmboldt, J., et al. 2015,
JGRA, 107, 1309
IAUGA, 22, 53643
Kiss, L. L., Szatm´ary, K., Cadmus, Jr., R. R., & Mattei,
J. A. 1999, A&A, 346, 542
Knapp, G. R., Bowers, P. F., Young, K., & Phillips, T. G.
1995, ApJ, 455, 293
Konstantinova-Antova, R., Auri`ere, M., Charbonnel, C.,
et al. 2010, A&A, 524, A57
Konstantinova-Antova, R., Aur´ıere, M., Charbonnel, C.,
et al. 2013, BlgAJ, 19, 14
Lim, J., Carilli, C. L., White, S. M., Beasley, A. J., &
Smith, A. M. S., Collier Cameron, A., Greaves, J., et al.
Kunitomo, M., Ikoma, M., Sato, B., Katsuta, Y., & Ida, S.
Lazio, W., T. J., Farrell, W. M., Dietrick, J., et al. 2004,
2011, ApJ, 737, 66
ApJ, 612, 511
Lecavelier des Etangs, A., Sirothia, S. K., Gopal-Krishna, &
Zarka, P. 2013, A&A, 552, A65
Marson, R. G. 1998, Natur, 392, 575
Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer,
P., & Lissauer, J. J. 2007, ApJ, 655, 541
Mizutani, H., Yamamoto, T., & Fujimura, A. 1992, AdSpR,
Murphy, T., Bell, M. E., Kaplan, D. L., et al. 2015,
12, 265
MNRAS, 446, 2560
Mustill, A. J., & Villaver, E. 2012, ApJ, 761, 121
Newell, R. T., & Hjellming, R. M. 1982, ApJL, 263, L85
Nordhaus, J., & Spiegel, D. S. 2013, MNRAS, 432, 500
Nordhaus, J., Spiegel, D. S., Ibgui, L., Goodman, J., &
Burrows, A. 2010, MNRAS, 408, 631
Noyola, J. P., Satyal, S., & Musielak, Z. E. 2014, ApJ, 791,
25
O'Gorman, E., Harper, G. M., Brown, A., Drake, S., &
Richards, A. M. S. 2013, AJ, 146, 98
Parsons, A., Pober, J., McQuinn, M., Jacobs, D., &
Aguirre, J. 2012, ApJ, 753, 81
Petit, P., Auri`ere, M., Konstantinova-Antova, R., et al.
2013, in LNP, Vol. 857, LNP, ed. J.-P. Rozelot & C. .
Neiner, 231
Pober, J. C., Liu, A., Dillon, J. S., et al. 2014, ApJ, 782, 66
Reimers, D. 1975, MSRSL, 8, 369
Reiners, A., & Christensen, U. R. 2010, A&A, 522, A13
Rogers, A. E. E., & Bowman, J. D. 2008, AJ, 136, 641
Russell, C. T. 1978, Natur, 272, 147
Sano, Y. 1993, JGG, 45, 65
Schild, H. 1989, MNRAS, 240, 63
Schoier, F. L., & Olofsson, H. 2001, A&A, 368, 969
Schroder, K.-P., & Cuntz, M. 2005, ApJL, 630, L73
17
-- . 2007, A&A, 465, 593
Sirothia, S. K., Lecavelier des Etangs, A., Gopal-Krishna,
Kantharia, N. G., & Ishwar-Chandra, C. H. 2014, A&A,
562, A108
Skinner, C. J., Dougherty, S. M., Meixner, M., et al. 1997,
MNRAS, 288, 295
2009, MNRAS, 395, 335
Spiegel, D. S. 2012, ArXiv e-prints, arXiv:1208.2276
Spiegel, D. S., & Burrows, A. 2012, ApJ, 745, 174
-- . 2013, ApJ, 772, 76
Spiegel, D. S., & Madhusudhan, N. 2012, ApJ, 756, 132
Starchenko, S. V., & Jones, C. A. 2002, Icar, 157, 426
Stevens, I. R. 2005, MNRAS, 356, 1053
Stevenson, D. J. 1979, GApFD, 12, 139
Stil, J. M., Keller, B. W., George, S. J., & Taylor, A. R.
Stroe, A., Snellen, I. A. G., & Rottgering, H. J. A. 2012,
2014, ApJ, 787, 99
A&A, 546, A116
Stromgren, B. 1939, ApJ, 89, 526
Suzuki, T. K. 2007, ApJ, 659, 1592
Treumann, R. A. 2006, A&ARv, 13, 229
Tsvetkova, S., Petit, P., Auri`ere, M., et al. 2013, A&A, 556,
A43
van den Oord, G. H. J., & de Bruyn, A. G. 1994, A&A,
van Haarlem, M. P., Wise, M. W., Gunst, A. W., et al.
van Loon, J. T., Cioni, M.-R. L., Zijlstra, A. A., & Loup, C.
2005, A&A, 438, 273
Vassiliadis, E., & Wood, P. R. 1993, ApJ, 413, 641
Verner, D. A., Ferland, G. J., Korista, K. T., & Yakovlev,
D. G. 1996, ApJ, 465, 487
Wu, C. S., & Lee, L. C. 1979, ApJ, 230, 621
Zarka, P. 1998, JGR, 103, 20159
Zarka, P., Cecconi, B., & Kurth, W. S. 2004, JGRA, 109, 9
Zarka, P., Lazio, J., & Hallinan, G. 2015, Advancing
Astrophysics with the Square Kilometre Array
(AASKA14), 120
Zarka, P., Treumann, R. A., Ryabov, B. P., & Ryabov,
V. B. 2001, Ap&SS, 277, 293
Parsons, A. R., Backer, D. C., Foster, G. S., et al. 2010, AJ,
286, 181
Pequignot, D., Petitjean, P., & Boisson, C. 1991, A&A, 251,
2013, A&A, 556, A2
139, 1468
680
APPENDIX
A. PLANETARY OUTFLOW
One might anticipate that, when a planet's host star reaches the red giant stage, the planet's atmosphere
could start to escape due both to the increased atmospheric temperature (from increased stellar irradiation,
which could lead to temperatures (cid:38)1000 K) and to the UV/X-ray radiation produced by accretion from
stellar wind. In this section, we briefly estimate these effects on the circumplanetary configuration.
The conventional escape parameter is computed for RGHJs as the ratio of escape energy to thermal
energy, which is the same as the squared ratio of escape speed to thermal speed:
(cid:18) Mp
(cid:19)(cid:18) Tc
(cid:19)−1(cid:18) rc
(cid:19)−1
MJ
1000K
RJ
λc =
GMpmp
kTcrc
= 220
(A1)
where Tc and rc are the radius and temperature at the exobase (the location above which the mean free
path of particles is longer than the pressure scale height). In this formalism, λc (cid:29) 1 implies Jeans escape,
whereas for λc approaching unity or below, the escape mechanism becomes a hydrodynamic flow. Eq. (A1)
suggests that, in most of the cases we consider, the atmospheric loss is in the Jeans regime. The rate of
Jeans escape is
ΦJeans[rc] =
1
σ
GMp
2R3
p
(λc + 1)
λce−λc
(A2)
where σ is the cross-section of hydrogen, here taken to be the geometric cross-section associated with the
Bohr radius. The ram pressure at distance r due to such an outflow is
(cid:115)
(cid:112)
(cid:18) r
(cid:19)−2
Rp
Pram[r] = mpΦJeans[rc]
vesc
(A3)
18
which indicates 2 × 10−7 dyne cm−2 even when λc = 1. In contrast, the ram pressure of the stellar wind
(the left side of equation 9) is ∼3×10−5 dyne cm−2, or two orders of magnitude greater than the outflow
ram pressure. The ram pressure of the planetary outflow would, therefore, be negligible.
B. IONIZATION CASCADE
For an electron-hydrogen ionizing collision, we estimate the differential cross-section σ as a function of
recoil energy dσ/dErecoil. We approximate the bound electron as a free electron at rest and use the differential
cross-section for the scattering of two electrons (Møller scattering) in the non-relativistic limit:
dσ
dΩCM
∼ 2c2α2
m2
ev2
rel sin4 θ
,
(B1)
where α is the fine structure constant, vrel is the relative velocity, θ is the scattering angle in the center of
mass (CM) frame, and me is the electron mass. The recoil energy (that is, the energy transferred to the
electron at rest) is related to the scattering angle via
and hence
We therefore have
Erecoil =
mev2
rel (1 − cos θ) ,
1
4
dErecoil
dΩCM
= − 1
4
mev2
rel .
dσ
dErecoil
∼
2c2α2
relE2
mev2
recoil
,
(B2)
(B3)
(B4)
which is largest for small recoil energies. Of course, Erecoil must be larger than ERydberg for this approxima-
tion to be meaningful.
An objection to the above estimate is that the singularity in the Erecoil → 0 limit arises from the long-
range nature of Coulomb interactions, which, of course, is absent in the real problem. Far away, the ionizing
electron sees the dipole moment of the neutral atom. However, for energy transfer of the order of a Rydberg
cα/2a0 we can estimate the transfer momentum and see that it is larger than the inverse of the Bohr radius
a0. This implies that, in this range, the above estimate may actually be reliable. The transfer momentum is
q2 =
1
4
Comparison with (B2) shows
Now set Erecoil ∼ cα/a0 to get
where we used a0 = /αm.
m2v2
rel(sin2 θ + (1 − cos θ)2) = m2v2
rel sin2 θ
2
Erecoil =
q ∼
(cid:114) mα
a0
q2
2m
.
∼
a0
.
(B5)
(B6)
(B7)
C. BACK-REACTION OF A PLASMA FLOWING INTO A MAGNETIC FIELD
Given the high stellar wind density of evolved stars, one might anticipate that the large quantity of
plasma flowing into a region permeated by a magnetic field would, by spiraling around the field lines,
generate an opposing magnetic field that partially cancels the intrinsic planetary magnetic field. To estimate
the strength of this effect, consider a uniform flow of particles of charge e, mass m, and velocity v into
a region of magnetic field B, spiraling along the initial magnetic field. A single charged particle moving
perpendicular to the magnetic field moves in a circular orbit of radius
r =
mv⊥
eB
.
(C1)
which creates a magnetic dipole
µ =
The volume integral of the canceling magnetic field Bc generated by a single dipole µ is
πr2 =
1
2
erv⊥ =
mv2⊥
2B
.
2πr/v⊥
e
(cid:90)
d3rBc =
8π
3
µ .
19
(C2)
(C3)
Hence, for a number density n of these charged particles, the fraction of the canceling field to the
background magnetic field is
Bc
B
=
n(mv2⊥/2)
3B2/8π
<
ρK
3ρB
,
where ρK and ρB are the kinetic energy density and the magnetic energy density, respectively.
Just inside the stand-off point,
mpn[rmag]v2 <
B[rmag]2
2π
and thus ρK/ρB < 1. When the particle spirals into the planet, the magnetic pressures increases as
On the other hand, the kinetic energy of the moment increases
B2
8π
∝ 1
r6 .
mv2
2
∝ 1
r
(C4)
(C5)
(C6)
(C7)
Therefore, unless the density increases more drastically than 1/r5, the ratio Bc/B is significantly less than
unity.
D. EXPECTED NUMBER OF DETECTIONS TAKING ACCOUNT OF PLANET
OCCURRENCE RATE
In order to estimate the expected number of RGHJs that might be detectable via near-future radio
observatories, we make use of the observed occurrence rate (around main-sequence stars) of Jovian planets.
We consider an empirical power-law planet occurrence rate described as
f [Mp, P ]≡
d2N
d log Mpd log P
(cid:18) Mp
(cid:19)α(cid:18) P
=C
MJ
1 day
(cid:19)β
,
(D1)
(D2)
where Cumming et al. (2008) found α = −0.31 ± 0.2, β = 0.26 ± 0.1, C = 1.04 × 10−3 as best-fit values.
Using this equation, the probability that a given evolved star in Figure 5 has an RGHJ with mass larger
than threshold Mp0 whose radio emission may be detectable is
(cid:90)
P[> Mp0] ∼
(cid:90) Pmax[Mp]
d log Mp
d log P f [Mp, P ] ,
(D3)
Mp0
Pmin[Mp]
where Pmin[Mp] is the minimum orbital period to have a cyclotron frequency larger than the ambient plasma
frequency, while Pmax[Mp] is the maximum orbital period, given the planetary mass, to emit detectable
radio flux. We solve for these critical period values by first solving for the corresponding semimajor axis
a3{min,max}/GM(cid:63). For the amin,
values, and then transforming to period via Kepler's law, P{min,max} = 2π
(cid:113)
20
Figure 6. The expected number of radio detection of Jovian exoplanets above the given threshold mass.
.
(D4)
combining equations (1), (36), and (38) leads to
amin[Mp]∼ 2.1 AU
For amax, equations (36) and (48) indicate
amax[Mp]∼ 3.8 AU
(cid:18) Mp
MJ
M(cid:63)
(cid:32)
×
10−8M(cid:12)/yr
t
4.5 Gyr
30 km/s
(cid:19)0.43
(cid:19)−1/2
(cid:19)−3/2
(cid:19)−1.04(cid:18)
(cid:33)1/2(cid:18) vsw
(cid:19)−3/4(cid:18) d
(cid:19)0.43/4
(cid:33)1/2(cid:18) v(cid:63)
(cid:19)5/4
4.5 Gyr
100 pc
t
(cid:18) Fmin
(cid:19)−1.04/4(cid:18)
10 µJy
M(cid:63)
10−8M(cid:12)/yr
10−1v(cid:12)
(cid:18) Mp
(cid:32)
MJ
×
×
.
(D5)
emission is (cid:80)
Using equation (D3), the expected number of planets more massive than Mp0 with detectable radio
i Pi[> Mp0], where Pi[> Mp0] is the existence and detection probability for the i-th target.
Figure 6 shows the expected number of detections as a function of threshold planetary mass Mp0. This figure
indicates that, if all of the targets in Figure 5 are RGB stars, the total number of detections would be about
four (seven), provided that the sensitivity of the radio instruments Fmin is 10 µJy (5 µJy). However, such
high sensitivity of the instrument is unlikely to be achieved at low frequency, so more massive planets are
probably the best candidates. If νcyc > 200 MHz (i.e., roughly Mp > 7MJ) is required for detectability, then
the expected number of detectable systems would be reduced to approximately one to two.
100101threshold planetary mass Mp0 [MJ]012345678cumulative expected numberFmin=10µJyFmin=5µJy |
1011.0186 | 2 | 1011 | 2011-04-29T20:05:58 | Aliases of the first eccentric harmonic : Is GJ 581g a genuine planet candidate? | [
"astro-ph.EP",
"astro-ph.IM",
"physics.data-an"
] | The radial velocity (RV) method for detecting extrasolar planets has been the most successful to date. The RV signal imprinted by a few Earth-mass planet around a cool star is at the limit of the typical single measurement uncertainty obtained using state-of-the-art spectrographs. This requires relying on statistics in order to unearth signals buried below noise. Artifacts introduced by observing cadences can produce spurious signals or mask genuine planets that should be easily detected otherwise. Here we discuss a particularly confusing statistical degeneracy resulting from the yearly aliasing of the first eccentric harmonic of an already-detected planet. This problem came sharply into focus after the recent announcement of the detection of a 3.1 Earth mass planet candidate in the habitable zone of the nearby low mass star GJ 581. The orbital period of the new candidate planet (GJ 581g) corresponds to an alias of the first eccentric harmonic of a previously reported planet, GJ 581d. Although the star is stable, the combination of the observing cadence and the presence of multiple planets can cause period misinterpretations. In this work, we determine whether the detection of GJ 581g is justified given this degeneracy. We also discuss the implications of our analysis for the recent Bayesian studies of the same data set, which failed to confirm the existence of the new planet. Performing a number of statistical tests, we show that, despite some caveats, the existence of GJ 581g remains the most likely orbital solution to the currently available RV data. | astro-ph.EP | astro-ph | Aliasing of the first eccentric harmonic : Is GJ 581g a genuine
planet candidate?
Guillem Anglada-Escud´e1 & Rebekah I. Dawson2
[email protected], [email protected]
ABSTRACT
The radial velocity (RV) method for detecting extrasolar planets has been the
most successful to date. The RV signal imprinted by a few Earth-mass planet
around a cool star is at the limit of the typical single measurement uncertainty
obtained using state-of-the-art spectrographs. This requires relying on statistics
in order to unearth signals buried below noise. Artifacts introduced by observing
cadences can produce spurious signals or mask genuine planets that should be
easily detected otherwise. Here we discuss a particularly confusing statistical
degeneracy resulting from the yearly aliasing of the first eccentric harmonic of an
already-detected planet. This problem came sharply into focus after the recent
announcement of the detection of a 3.1 Earth mass planet candidate in the hab-
itable zone of the nearby low mass star GJ 581. The orbital period of the new
candidate planet (GJ 581g) corresponds to an alias of the first eccentric harmonic
of a previously reported planet, GJ 581d. Although the star is stable, the com-
bination of the observing cadence and the presence of multiple planets can cause
period misinterpretations. In this work, we determine whether the detection of
GJ 581g is justified given this degeneracy. We also discuss the implications of
our analysis for the recent Bayesian studies of the same dataset, which failed to
confirm the existence of the new planet. Performing a number of statistical tests,
we show that, despite some caveats, the existence of GJ 581g remains the most
likely orbital solution to the currently available RV data.
1
1
0
2
r
p
A
9
2
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
6
8
1
0
.
1
1
0
1
:
v
i
X
r
a
Subject headings:
niques: radial velocities -- Methods: data analysis
Stars:
individual (GJ 581) -- planetary systems -- Tech-
1Carnegie Institution of Washington, Department of Terrestrial Magnetism, 5241 Broad Branch Rd. NW,
Washington D.C., 20015, USA
2Harvard-Smithsonian Center for Astrophysics, 60 Garden St, MS-10, Cambridge, MA 02138
-- 2 --
1.
Introduction
The recently reported planet candidate around the nearby M dwarf GJ 581 (Vogt et al.
2010, hereafter V10) has generated much public enthusiasm and a similar amount of skep-
ticism within part of the scientific community (Pepe 2010).
If confirmed, it will be the
first planet potentially capable of hosting life as we know it now (Kasting et al. 1993;
von Braun et al. 2011; Heng & Vogt 2010). The possible existence of this planet was an-
nounced based on analysis of the new HIRES/Keck precision RV measurements combined
with HARPS/ESO data published by Mayor et al. (2009) (hereafter M09). V10 reported
that the candidate planet GJ 581g (hereafter planet g) has a minimum mass of 3.1 M⊕ and
a period of 36.5 days. The data for GJ 581 contain the signal of at least 4 other low mass
planets (planets b − e). Dynamical studies showed that their masses cannot be larger than
∼1.4 times the reported minimum masses (M09 and V10).
Of particular interest to this work is GJ 581d. The planet has a period of 67 days
and a previously reported eccentricity of ∼ 0.4 (M09). Using only the HARPS data (M09),
Anglada-Escud´e et al. (2010a) noted that the eccentric orbit of GJ 581d could disguise the
signal of a low mass companion at half its period, similar to the candidate planet g now
reported by V10. Both eccentric planets and pairs in or near 2:1 mean-motion resonance have
been discovered by RV and transit surveys1. Each case implies a very different dynamical
history so the distinction is important in understanding evolution of planetary systems.
The available RV data of GJ 581 is strongly affected by yearly aliasing (Dawson & Fabrycky
2010, hereafter DF10). In Section 2, we explain this aliasing and demonstrate how the ec-
centricity of a planet can be confused with an additional planet at half its period. In Section
3, we show that planet GJ 581g has an orbital frequency that aliases to the eccentric har-
monic of planet d and therefore could be an artifact of the sampling cadence. In Section
4, we reevaluate the significance of planet g in the presence of this degeneracy by allowing
all the planetary orbits to be eccentric. In the same section, we quantify the probability of
confusing planet g with the eccentric harmonic of planet d. In our analysis, we encountered
some caveats about the reality and uniqueness of the reported signal. In Section 4.4, we
describe these caveats and their relation with the failure of recent Bayesian studies to con-
firm the presence of planet g (Tuomi 2011; Gregory 2011). Given that the 6-planet solution
proposed by V10 is the most significant one that is also physically allowed, we conclude that
the presence of planet g remains supported by the data.
1see exoplanet.eu for an up-to-date census
-- 3 --
2. Sampling cadence and aliases
The sampling cadence of a periodic signal can alter its apparent period. An alias is a
spurious periodicity generated by finite sampling of a real signal. As an extreme example, if
a function of period P is measured at regular intervals ∆t = P , it will appear constant to the
observer. In the case of even sampling, the maximum frequency that can be probed without
ambiguity is half the sampling frequency (the Nyquist frequency). However, this frequency
cannot be unambiguously determined in the case of uneven sampling (Eyer & Bartholdi
1999). Instead, one can identify characteristic sampling frequencies by computing the spectral
window function W , which depends only on the observation instants. The modulus of W
ranges from 0 to 1 and peaks at the sampling frequencies that cause the most severe aliases.
For a real signal of frequency f = 1/P and a characteristic sampling frequency fs, aliases
appear at f ± fs. Figure 1 shows the window function including all the HARPS+HIRES
data available for GJ 581 (M09 and V10). Optical astronomical measurements occur only
at night and only during the season when the star is observable. The night time cadence
introduces two sampling frequencies close to 1 day−1 (the solar and sidereal day, top right
panel), and the seasonal availability generates a 1 year peak (top left panel). When the
signal -- to -- noise is low, coherent addition of noise can produce higher power at the aliases than
at the real period (see DF10). Since the announcement of the first planet candidates around
GJ 581 by Bonfils et al. (2005), the yearly alias has significantly affected the determination
of the orbital periods of these planets. For example, the first period proposed for the super-
Earth GJ 581d was 84 days (Udry et al. 2007), a yearly alias of the currently accepted period
of 67 days.
The other ingredient to the degeneracy is that the RV signal due to a planet's Keplerian
motion can be expressed as a series in powers of its orbital eccentricity, e. To first order in
eccentricity, a Keplerian RV signal of amplitude K and period P can be written as
vr(t) = γI + K cos(cid:18) 2πt
P
+ φ0(cid:19) + Ke cos(cid:18) 2πt
P/2
+ φ1(cid:19) + O(Ke2) ,
(1)
where φ0 and φ1 are functions of the initial mean anomaly and the argument of the periastron,
and γI is a constant offset that can be different for each instrument. Eqn. 1 implies that
in a system with a planet of period P , an additional body with period P/2 is statistically
indistinguishable unless the next term in the expansion can be measured. We call the term
proportional to Ke the "first eccentric harmonic." Due to statistical uncertainties, finite
sampling, and aliasing, two planets can be misinterpreted as a single eccentric planet even
when their period ratio is not exactly 2.
-- 4 --
3. Degeneracy between planets GJ 581d and g
As shown by M09, the period of planet d is ∼ 67 days. Therefore, the first eccentric
harmonic of this planet has a period of ∼ 33.5 days (f = 0.02985 day−1). The period of
the newly found planet g is 36.5 days, and its aliases should appear at 33.18 and 40.55 days
(f = 0.02985 day−1±1/yr). Because a yearly alias of planet g falls very close to the eccentric
harmonic of planet d, planet g's signal will be partially absorbed by the eccentricity of planet
d. To illustrate this, we carry out the following experiment. We generate a synthetic dataset,
the exact 6-planet solution from V10 (all circular orbits) sampled at the times of the HARPS
and HIRES datasets and with Gaussian noise added (1.9 and 2.7 m/s respectively). Then,
we compute the periodogram of the residuals of a 4-planet fit to this synthetic dataset (Fig.
2). Even though planet g is included in the simulated data, its peak will disappear if planet
d is allowed to be eccentric (top left). On the other hand, planet g will appear prominently
if the eccentricity of planet d is preliminarily fixed at 0 (bottom left). The same behavior is
observed for the real data (right panels). As a general rule in modeling RV data sets, it is
safer to search for significant periodicities first (i.e. preliminarily fixing the eccentricities at
0) and defer determining which model is preferred to a later stage (i.e. an eccentric planet
vs. a two planet model).
In DF10, a test was developed to qualitatively assess which period is more likely in the
presence of signal aliasing. The test consists of the following steps: (1) generate noiseless
synthetic datasets of the signals under study, (2) compute their power spectra at the regions
where the most prominent features occur (the nominal period and strongest aliases), and
(3) compare them with the power spectrum of the real data (i.e. for GJ 581, the residuals
to the 4-planet fit with eccentricities preliminarily fixed at 0). By power spectrum we mean
the amplitude of the sinusoidal function that best fits the data at each frequency (see DF10
for further details).
We generate a synthetic dataset for each of the two candidate periods (33.5 days and
36.5 days). The bottom panels of Figure 1 show the power spectra of the real and the
synthetic datasets at the nominal periods and their most prominent aliases. According to
Lomb (1976), "If there is a satisfactory match between an observed spectrum and a noise-
free spectrum of period P , then P is the true period". While the eccentric harmonic (33.5
days) and its aliases fail to mimic the features observed in the real data (especially the daily
aliases in the central panels), the candidate signal at 36.5 days does a fairly good job of
reproducing most of them. Random noise modifies the power balance between peaks, so one
should understand this as a first qualitative assessment. The probability of confusing planet
g with the eccentric harmonic of planet d is quantified in Section 4.3.
-- 5 --
4. Detailed analysis
4.1. Preliminary period search
In order to indentify promising low-amplitide periodicities, first we sequentially subtract
the four most significant signals that coincide with those reported by M09 (using the sys-
temic interface (Meschiari et al. 2009)). As discussed in Section 3, we preliminarily fix the
eccentricities at 0. All parameters are refined each time a planet is added. The periodogram
(Anglada-Escud´e et al. 2010b) of the residuals to the 4-planet model is then computed. Both
the 36.5 and 433 day signals appear prominently, see bottom right panel of Fig.2. A detection
False Alarm Probability (FAP) is obtained through a Monte Carlo approach using 104 syn-
thetic realizations of the data. Each synthetic dataset is created by scrambling the residuals
to the four -- planet fit, keeping the same observing epochs and instrument membership. The
detection FAP is the fraction of times we obtain a signal with a higher power than the one in
the real residuals. We find that both detection FAP are below 0.5%. After subtracting these
two signals (36.5 and 433 days), no additional peaks are found with a detection FAP under
1%. These FAP are for detection of sinusoid -- like signals only. The actual significances for
the proposed new candidates are computed in the next section.
4.2. Statistical significance of GJ 581g
Recently, Gregory (2011) and Tuomi (2011) indicated that planet g is not robustly
confirmed when Bayesian analysis is applied. To address this issue, we consider in this
section the general case where the orbits of the already-detected planets are allowed to be
eccentric and compute the definitive FAP analytically using a classical Frequentist approach.
Because the uncertainties in RV measurements are difficult to quantify, the χ2 statistic cannot
be used to evaluate the goodness-of-fit in an absolute sense (Andrae et al. 2010). However,
the χ2 can still be used in a differential way, i.e., to determine whether or not the addition of
i (ri are the residuals with
respect to a proposed model and ǫi are the uncertainties). In this respect, we apply a version
of the Fisher F-ratio test proposed by Cumming et al. (2008) that allows the addition of
any number of free parameters to the null hypothesis and to the model being tested. This
requires the computation of the F-ratio as
a planet is justified given the improvement of the χ2 = Pobs
r2
i /ǫ2
i
F =
(χ2
null − χ2
+)/(k+ − knull)
χ2
+/(Nobs − k+)
(2)
-- 6 --
where knull and k+ are the number of free parameters of the null hypothesis and the model
to be tested respectively. The F -- ratio follows an F -- probability distribution with (k+ − knull)
and (Nobs − k+) degrees of freedom (Johnson et al. 1995). The null hypothesis includes
the M09 planets (planets b, c, d, e) and planet f , all with fully Keplerian orbits. The χ2
+
is obtained by adding a new planet on a ∼36.5 day circular orbit (requiring two extra
parameters) and readjusting all the free parameters. Using the values in the two last columns
of Table 1, the obtained probability of improvement-by-chance at that particular frequency
is p = 4.41 × 10−5. Now one has to consider the number M of independent frequencies
that could also generate a spurious improvement. According to V10, M is 2525 for the
HARPS+HIRES combined dataset. Therefore, the definitive FAP for planet g becomes
FAP= 1 − (1 − p)M (Cumming et al. 2008) and amounts to 0.11%. For completeness, we add
the FAP of planet f (∼ 0.03%) to Table 1, using the Keplerian 4-planet solution (M09) as
the null hypothesis.
The F-ratio test can also be used to evaluate the significance of the eccentricity compared
to a circular orbit (the null hypothesis). We find that the eccentricities of the larger amplitude
planets (i.e. b, c, d and e) are well-constrained but compatible with 0. For the low amplitude
candidates (f and g), the eccentricities are unconstrained. Table 2 provides the full orbital
solution.
4.3. Probability of confusion
At this point, we have assessed the FAP of the planet detection. Now we need to quantify
the probability that the eccentric harmonic of planet d could inflate the signal at 36.5 days
by an unfortunate combination of random errors. We generate 104 synthetic datasets by
scrambling the residuals to the 6-planet fit and injecting a signal at 33.5 days (amplitude
1.29 m/s). The periodogram of each dataset is then computed. We define the probability of
confusion as the number of times the period of the highest peak does not lie within 1 day
of the injected signal, divided by the total number of trials. For the 33.5 days, we find this
probability is 0.34%. In 0.12% of trials, the highest peak is found near the 36.5 days alias.
Performing the same test by injecting the 36.5 days signal, we obtain a 0.5% probability of
confusion (with 0.2% of trials corresponding to the 33.5 days alias). As a final check, we
repeat the same experiment using random noise instead of scrambling the residuals (σ = 1.8
m/s for HARPS, and σ = 3.0 m/s for HIRES). Similarly, low confusion rates (∼ 0.5%) are
obtained. In every case, the probabilities are below 1%, supporting the conclusion of the
previous section: the probability of confusing planet g with the eccentric harmonic of planet
d is very low. As a final check, we apply the same test using only the HARPS residuals. In
-- 7 --
more than 50% of the trials, the highest peak is none of the injected signals, confirming that
the data in M09 was not sensitive enough to distinguish between cases.
4.4. Caveats
During our analysis, we encountered a number of caveats that will require further in-
vestigation. Recent analyses by other authors using Bayesian methods (i.e. Gregory 2011;
Tuomi 2011) seem to contradict the conclusions reached here. Thus we discuss some of these
caveats and possible lines of inquiry.
Significance of spurious periodicities. We found a number of alternative periodic-
ities for planets f and g that yield alternative 6-planet fits with similar significance to the
solution announced by V10. Potential periods for planets f and g include 25.0, 26.8, 30.6,
47.8, 54.6, 59.3, 71.4, 76.4 and 97.9 days. These alternative solutions were found by comput-
ing a two-dimensional periodogram (i.e. two periodic signals are adjusted simultaneously on
a grid) of the residuals of the 4-planet circular solution and refining all Keplerian parameters
around the areas of lowest χ2 (Markwardt 2009). However, a case-by-case investigation in-
dicated that all correspond to orbital configurations with high eccentricities that suffer from
several orbital crossings, making them dynamically unstable on very short timescales. To
ultimately decide if such solutions were acceptable, we fixed the involved eccentricities to
lower values, adjusting all other parameters in the process. The resulting orbital fits were
poorer than the 6-planet solution from V10 (even assuming circular orbits for planets f and
g). The significance of these unphysical solutions raises the caveat that planet g could be
a physically-possible but spurious signal. As discussed by Tuomi (2011), Bayesian analysis
methods do not yet apply down-weighting of unphysical configurations. Because such orbits
have high significance, they will be oversampled, downgrading the likelihood of physically
allowed orbits. This may explain the apparent contradiction of the recent Bayesian analysis
with our conclusions.
Hidden planets. While planets b and c are detectable in both datasets, planets d
and e are not obvious in the HIRES data alone (as also noted by V10). This indicates that
some signals may be consistent with a given dataset, yet not independently detectable due
to sampling issues. As a similar case, GJ 876e was a strong signal in the HIRES dataset yet
undetectable with HARPS (Rivera et al. 2010). Because the completeness of on-going RV
surveys can be affected as a result, this caveat requires further investigation.
Noise unknowns. We have chosen statistical tools that are as insensitive as possible to
assumptions about noise: the F-ratio test and Monte Carlo scrambling of the residuals. Still,
-- 8 --
Table 1. Statistical quantities required to evaluate the F-ratio in Eqn. 2
Parameter
Four planets Four planets
M09 solution + circular f
Five planets Five planets
+ circular g
Planets includeda
Eccentric orbits
Circular orbits
Eccentricity
of planet d
RMS (m/s)
Free parameters
Nobs
χ2
F ratio
FAP
bced
bcde(f)
bcdef
bcdef(g)
4
0
0.5
2.32
22
241
707.7
4
1
0.4
2.21
24
241
611.8
17.0
0.03%
5
0
0.4
2.20
27
241
602.5
5
1
∼ 0.1
2.10
29
241
524.8
15.69
0.11%
aA planet name in parentheses indicates that this solution was used to compute the FAP
of that planet. The number of free parameters is obtained as follows: 2 RV offsets (one per
instrument), 5 for each Keplerian planet, and 2 for the circular orbit to be tested.
-- 9 --
we recognize that it is not fully understood how systematic effects (e.g. stellar jitter) impact
the sensitivity to low amplitude signals. A Bayesian approach with a noise parameter has
been proposed to solve this problem (Tuomi 2011). However, we think that further tests are
required to assess the sensitivity of Bayesian methods in the presence of unknown systematic
noise and low amplitude signals.
5. Conclusions
The eccentric harmonic of planet d coincides with a yearly alias of the newly reported
planet g, meaning that both signals are correlated and that a premature Keplerian fit to
planet d prevents the detection of planet g. We have found a number of unphysical solutions
that fit the data just as well. Still, the proposed pair of planets f and g remains the only
physically viable solution that significantly improves the 4-planet fit. Thus we are compelled
to conclude that the presence of GJ 581g is well supported by the available data. The
ultimate confirmation will require additional RV measurements and reanalysis of the data
in a very convincing way. To mitigate yearly aliasing, we encourage observations at more
extreme parallax factors.
We end with two cautionary notes. First, whether or not planet g exists, the same
cadence issues may be present in other datasets. We have found that significant periodicities
in one dataset do not appear in another and that several unphysical models can appear
significant. If future observations rule out the existence of planet g, the fact that it passes
the standard, widely-used statistical tests would bode ill for other low-amplitude planet
candidates. We remark that statistical significance tests can be very sensitive to assumptions
by the authors (including this work). Bayesian methods may provide stricter confidence level
estimates but additional testing is required to ensure that they are not over-conservative in
the low signal -- to -- noise regime (Jenkins & Peacock 2011). Bayesian methods may also need
to demonstrate how their conclusions are impacted by sampling issues and the inclusion of
dynamical constrains in the likelihood sampling process.
Second, the eccentricity of a long period giant planet may cause spurious low amplitude
signals mimicking habitable planets around Sun-like stars. Consider Jupiter as an example:
its eccentric harmonic would have a period of 5.9 years (or f = 0.00046162 days−1). The
corresponding yearly aliases would be at f ±0.0027378 days −1, giving candidate signals at
439 and 313 days. Complementarily, a genuine planet candidate can be missed if the eccen-
tricity of an outer planet is adjusted prematurely. Aliasing tests on the eccentric harmonics
of detected planets might be necessary in future claims of very low-amplitude signals.
-- 10 --
Acknowledgements This work was funded by the Carnegie Fellowship Postdoctoral
Program and the National Science Foundation Graduate Research Fellowship. We thank
P. Butler, and S. Vogt for early access to the RV data. We thank A. Boss, J. Chanam´e,
J. Dunlap, D. Fabrycky, N. Haghighipour, M. Lopez-Morales, and N. Moskovitz for useful
comments and discussions.
REFERENCES
Andrae, R., Schulze-Hartung, T., & Melchior, P. 2010, ArXiv e-prints
Anglada-Escud´e, G., L´opez-Morales, M., & Chambers, J. E. 2010a, ApJ, 709, 168
Anglada-Escud´e, G., Shkolnik, E. L., Weinberger, A. J., Thompson, I. B., Osip, D. J., &
Debes, J. H. 2010b, ApJ, 711, L24
Bonfils, X., et al. 2005, A&A, 443, L15
Cumming, A., Butler, R. P., Marcy, G. W., Vogt, S. S., Wright, J. T., & Fischer, D. A.
2008, PASP, 120, 531
Dawson, R. I., & Fabrycky, D. C. 2010, ApJ, 722, 937
Eyer, L., & Bartholdi, P. 1999, A&AS, 135, 1
Gregory, P. C. 2011, ArXiv e-prints 1101.0800
Heng, K., & Vogt, S. S. 2010, ArXiv e-prints 1010.4719
Johnson, N. L., Kotz, S. & Balakrishnan N. 1995, Continuous Univariate Distributions, Vol.
2. ed. Wiley (Sec. Edition, Sec. 27)
Jenkins, C. R., & Peacock, J. A. 2011, MNRAS, 338
Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993, Icarus, 101, 108
Lomb, N. R. 1976, Ap&SS, 39, 447
Markwardt, C. B. 2009, in Astronomical Society of the Pacific Conference Series, Vol. 411,
Astronomical Society of the Pacific Conference Series, ed. D. A. Bohlender, D. Du-
rand, & P. Dowler, 251 -- +
Mayor, M., et al. 2009, A&A, 507, 487
-- 11 --
Meschiari, S., Wolf, A. S., Rivera, E., Laughlin, G., Vogt, S., & Butler, P. 2009, PASP, 121,
1016
Pepe, F. 2010, IAU Symposium : The Astrophysics of Planetary Systems. Oral Contribution
Rivera, E. J., Laughlin, G., Butler, R. P., Vogt, S. S., Haghighipour, N., & Meschiari, S.
2010, ApJ, 719, 890
Tuomi, M. 2011, A&A, 528, L5+
Udry, S., et al. 2007, A&A, 469, L43
Vogt, S. S., Butler, R. P., Rivera, E. J., Haghighipour, N., Henry, G. W., & Williamson,
M. H. 2010, ApJ, 723, 954
von Braun, K., et al. 2011, ApJ, 729, L26+
This preprint was prepared with the AAS LATEX macros v5.2.
-- 12 --
Table 2. Keplerian solution including 6 planets (V10). Parameter values correspond to
the best χ2 solution. Uncertainties have been obtained using the Markov Chain Monte
Carlo algorithm included in the systemic interface (Meschiari et al. 2009).
Parameter
b
c
d
e
P [days]
e
K [ms−1]
ω [deg]
M0 [deg]
5.368483 (10−5)
10−4(10−3)(∗)
12.49 (0.15)
87.9 (10)
190.3 (8)
12.91750 (0.002)
0.10 (0.05)(∗)
3.40 (0.18)
190.6 (7)
193.3 (5)
66.845 (0.09)
0.11 (0.08)(∗)
1.90 (0.15)
61.5 (8)
353.9 (7)
3.1485 (2 10−4)
0.18 (0.08)(∗)
1.75 (0.14)
104.2 (7)
144.7 (5)
mp sin i [M⊕]
a [AU]
15.61 (0.16)
0.041
5.71 (0.23)
0.073
5.4 (0.21)
0.218
1.805 (0.16)
0.028
New planets (V10)
f
g
P [days]
e
K [ms−1]
ω [deg]
M0 [deg]
mp sin i [M⊕]
a [AU]
γHARPS [ms−1]
γHIRES [ms−1]
χ2
RMS [m/s]
36.53 (0.32)
0.19(u)
1.34 (0.17)
68.9(u)
248.1 (10)
3.18 (0.4)
0.146
440.7 (2.0)
0.17(u)
1.18 (0.13)
330.8(u)
183.24 (10)
6.28 (0.65)
0.767
-2.22 (0.2)
0.934 (0.2)
520.5
2.093
(u)Unconstrained parameter
(∗)Compatible with 0
-- 13 --
Period (days)
P = 0.997 day
Window function
P = 1 year
W
Frequency (1/days)
Period (days)
Eccentric harmonic
Planet g
Real residuals
)
s
/
m
(
K
Frequency (1/days)
Fig. 1. -- Upper panel: Modulus of the spectral window function including HARPS (M09)
and HIRES (V10) data. The top right panel shows both the sidereal and solar day as
significant sampling frequencies. The top left panel shows that the 1 year−1 frequency will
also create aliases. Bottom panel: Power spectra of the residuals to a 4-planet fit with
eccentricities preliminarily fixed at 0 (bottom), a noiseless synthetic planet g (center), and
a noiseless synthetic signal at the frequency of the eccentric harmonic (top). The aliases are
marked with symbols corresponding to the upper panel.
-- 14 --
10
8
6
4
2
0
10
8
6
4
2
0
Simulated data
?
planet f
25
125
625
planet g
planet f
25
125
Period (days)
625
16
12
8
4
0
16
12
8
4
0
Real data
?
planet f
25
125
625
planet g
planet f
25
125
Period (days)
625
r
e
w
o
P
r
e
w
o
P
d
t
e
n
a
p
l
c
i
r
t
n
e
c
c
E
d
t
e
n
a
p
l
l
r
a
u
c
r
i
C
Fig. 2. -- Periodograms of the residuals to a 4-planet fit containing the planets in M09.
Top: periodograms assuming an eccentric planet d. Bottom: periodograms of the same
datasets, fixing ed at 0. The left panels correspond to synthetic data containg the six planets
announced by V10. The right panels are periodograms of the residuals of actual data. The
36.5 day signal will be missed if the eccentricity of planet d is prematurely adjusted (red
question marks).
|
1703.03759 | 2 | 1703 | 2017-05-10T14:45:49 | Orbital Evolution of Moons in Weakly Accreting Circumplanetary Disks | [
"astro-ph.EP"
] | We investigate the formation of hot and massive circumplanetary disks (CPDs) and the orbital evolution of satellites formed in these disks. Because of the comparatively small size-scale of the sub-disk, quick magnetic diffusion prevents the magnetorotational instability (MRI) from being well-developed at ionization levels that would allow MRI in the parent protoplanetary disk. In the absence of significant angular momentum transport, continuous mass supply from the parental protoplanetary disk leads to the formation of a massive CPD. We have developed an evolutionary model for this scenario and have estimated the orbital evolution of satellites within the disk. We find, in a certain temperature range, that inward migration of a satellite can be stopped by a change in the structure due to the opacity transitions. Moreover, by capturing second and third migrating satellites in mean motion resonances, a compact system in Laplace resonance can be formed in our disk models. | astro-ph.EP | astro-ph |
Draft version April 27, 2018
Preprint typeset using LATEX style emulateapj v. 8/13/10
ORBITAL EVOLUTION OF MOONS IN WEAKLY ACCRETING CIRCUMPLANETARY DISKS
Yuri I. Fujii1, Hiroshi Kobayashi2, Sanemichi Z. Takahashi3, Oliver Gressel1
1 Niels Bohr International Academy, The Niels Bohr Institute, Blegdamsvej 17, DK-2100, Copenhagen Ø, Denmark
2 Department of Physics, Nagoya University, Furo-cho, Showa-ku, Nagoya, Aichi, 464-8602
3 Astronomical Institute, Tohoku University, 6-3 Aramaki, Aoba-ku, Sendai, Japan, 980-8578
Draft version April 27, 2018
ABSTRACT
We investigate the formation of hot and massive circumplanetary disks (CPDs) and the orbital evolution of satellites
formed in these disks. Because of the comparatively small size-scale of the sub-disk, quick magnetic diffusion prevents
the magnetorotational instability (MRI) from being well-developed at ionization levels that would allow MRI in the
parent protoplanetary disk. In the absence of significant angular momentum transport, continuous mass supply from
the parental protoplanetary disk leads to the formation of a massive CPD. We have developed an evolutionary model
for this scenario and have estimated the orbital evolution of satellites within the disk. We find, in a certain temperature
range, that inward migration of a satellite can be stopped by a change in the structure due to the opacity transitions.
Moreover, by capturing second and third migrating satellites in mean motion resonances, a compact system in Laplace
resonance can be formed in our disk models.
Subject headings: planets and satellites: formation – planets and satellites: gaseous planets – proto-
planetary disks
1. INTRODUCTION
Regular satellites around gas-giant planets are thought
to form in a surrounding gaseous disk. This notion is sup-
ported by the near-circular orbits of the moon systems
in our own Solar System, which are well-aligned to their
equatorial-planes, except for irregular satellites that have
been captured dynamically.
In analogy with a minimum mass solar nebula (Hayashi
1981) for planet formation, minimum mass sub-nebula
models were introduced for satellite formation (e.g.
Lunine & Stevenson 1982). For explaining certain char-
acteristics of the Galilean moons, Canup & Ward (2002,
2006) introduced the so-called gas-starved disk model
and reproduced the ratio between the planet mass and
the total mass of the satellite system. Satellite formation
around gas giants in a solids-enhanced minimum mass
model was discussed by Mosqueira & Estrada (2003a,b)
and Estrada & Mosqueira (2006) moreover developed a
scenario in gas-poor environment. Sasaki et al. (2010) in-
troduced an inner cavity in a gas-starved disk and found
that several moons in a resonance can be formed. Based
on this work, Ogihara & Ida (2012) performed N-body
simulations and have shown that moons are commonly
captured in a 2:1 mean motion resonance outside the
cavity and Galilean-like configuration can be formed.
In contrast
to the nebula hypothesis,
relatively
small rocky satellites present today can be well ex-
plained by a formation scenario in a tidal spread-
ing disk (Charnoz et al. 2010; Crida & Charnoz 2012;
Hyodo et al. 2015). The model is compelling, but it re-
quires a disk of solid material as the starting point. Such
circumplanetary "debris" disks may originate from the
capture of planetesimals (e.g. Hyodo et al. 2016, 2017)
or from tidal disruption of a previous generation of satel-
lites. However, larger satellites, especially those that
maintain an atmosphere, need gas around them during
their accretion. Therefore, it is reasonable to assume that
[email protected] (YIF)
at least some of the regular satellite systems must have
originated from gaseous circumplanetary disks (CPDs).
When a protoplanet grows to the size of several
earth masses in a protoplanetary disk (PPD) compa-
rable to the one typically assumed for the early solar
nebula (Hayashi 1981), gas around the planet starts
to accrete onto it. At that time, because of the con-
servation of angular momentum, a rotationally sup-
ported disk forms around the planet. On theoreti-
cal grounds, CPDs can be observed in many hydrody-
namic and magnetohydrodynamic (MHD) simulations
(e.g. Tanigawa & Watanabe 2002; Klahr & Kley 2006;
Ayliffe & Bate 2009; Machida et al. 2010; Szul´agyi et al.
2014, 2017; Gressel et al. 2013; Perez et al. 2015). Be-
cause of the orders of magnitude difference in spatial
scales, however, resolving the very vicinity of the planet
in those simulations is still difficult. Tanigawa et al.
(2012) have successfully measured the mass infall rate
onto a CPD during the early stage of its evolution, but
the long-term evolution remains to be established. Yet,
the modeling a CPD fully self-consistently during the
full PPD and planetary gap evolution is difficult just like
modeling the formation of PPDs from cloud-collapse is
difficult.
In PPDs, the magnetorotational instability (MRI) is
thought to play an important role in facilitating the ac-
cretion of the disk gas. Although angular momentum
transfer in a CPD was previously expected to be as ef-
fective as that in a PPD, sustaining the MRI is more diffi-
cult in CPDs (Fujii et al. 2011, 2014; Turner et al. 2014;
Keith & Wardle 2014). Thermal ionization can trigger
MRI at the inner radii of a CPD if the temperature be-
comes sufficiently high (Keith & Wardle 2014). However,
in the absence of strong MRI turbulence, gas accretion
may not be efficient enough to prevent a CPD from be-
coming massive by accumulation of infalling material.
An alternative scenario for angular-momentum transport
within the CPD may be provided by a magnetocentrifu-
gal disk wind which has been found to operate sporadi-
2
Fujii, Kobayashi, Takahashi & Gressel
cally in resistive-MHD simulations (Gressel et al. 2013).
It remains to be shown whether CPD winds are equally
emerging when including additional micro-physics such
as ambipolar diffusion.
In any case, disk outflows are
generally competing with infall, and it is unclear how a
steady state can be reached for CPDs that are still deeply
embedded in their parent disks.
If the sub-disk grows so massive as to become gravita-
tionally unstable, spiral arms appear and transfer angu-
lar momentum. Whether or not the gravitational energy
is converted into heat in situ is still under debate, but
supposed the energy deposition is local, the tempera-
ture with the CPD can become high.
In such a situa-
tion, episodic accretion caused by a combination of the
gravitational instability (GI) and the MRI (boosted by
thermal ionization) is to be expected. Martin & Lubow
(2011, 2013) and Lubow & Martin (2012, 2013) studied
this phenomenon in a layered CPD model that is devel-
oped in the context of PPDs (e.g. Armitage et al. 2001).
As mentioned earlier, a CPD is likely to become mas-
sive in the absence of significant transport of angular
momentum, that is if the temperature is insufficient to
maintain MRI turbulence. In this paper, we develop an
alternative model of the satellite-forming region of CPDs
considering the mass inflow from the PPD as the dom-
inant factor. Because there still is a gap between the
spatial scales that can be well-resolved by full-blown 3D
MHD simulations and the actual satellite-forming region
(inside a few tens of planet radii), we pursue the strategy
of 1D modeling of the sub-disk by means of an effective
description based on results from numerical simulations.
By doing so, in some of our models, we find a bump
in the radial surface density structure. We will exam-
ine whether such a specific location can stop the migra-
tion of moons. Given the situation that the innermost
moon survives rapid inward type-I migration by conver-
gent migration to the pressure maximum, we investigate
the possibility of trapping the second and third moons
in a mean motion resonance (MMR). The inner three of
the Galilean moons are known to be in a 4:2:1 mean mo-
tion resonance, a so-called Laplace resonance. In some of
our models, we successfully obtained a system in Laplace
resonance.
This paper is organized as follows. In Section 2, we de-
scribe our sub-disk model and assumptions and resulting
disks are shown in 3. We then highlight several models
with interesting structure and discuss the orbital evolu-
tion of moons in the disks in Section 4. Discussion of
the obtained results and a brief summary are given in
Sections 5 & 6, respectively.
2. MODELING OF CIRCUMPLANETARY DISKS
2.1. Derivation of Surface Density and Temperature
The equation for the time evolution of surface density
is essentially derived in the same way as in Fujii et al.
(2014), but in addition, here we simultaneously solve for
the temperature structure of the embedded sub-disk.
We determine the surface-density profile of the CPD
by solving a diffusion equation with an additional source
term stemming from mass infall from the parent PPD.
When the sub-disk's angular velocity is taken to be Ke-
plerian, the evolution of the surface density is given by
∂Σ
∂t
=
1
r
(1)
∂
∂r (cid:20)3r1/2 ∂
∂r (cid:16)r1/2νΣ(cid:17)(cid:21) + f ,
where r is the radius, ν is the kinematic viscosity
coefficient, and f is the mass flux from infall onto
the CPD. We employ the standard α prescription of
Shakura & Sunyaev (1973), namely, ν = αc2
s /ΩK with
cs being the sound speed and ΩK the Keplerian rotation
frequency.
To determine a prescription for the source term f , we
adopted the results of a detailed analysis of the 3D high-
resolution simulations by Tanigawa et al. (2012). Even
though they employed 11 levels of nested grids, the res-
olution is insufficient to resolve the vertical structure of
the CPD in the innermost several Jupiter radii from the
planet. Therefore, Tanigawa et al. measured physical
values of infalling material at high altitude, where the
infall is supersonic – the idea being that the infall rate
obtained in this way is not affected by the uncertainty
caused by the architecture of the CPD further down-
stream. The effective mass flux onto an inner part of
CPD is f (r) ∝ r−1 (see Figure 15 of Tanigawa et al.
2012). We assume that the planet has 0.4 MJ and lies
5.2 AU away from a solar mass star. Based on the val-
ues at this distance in the minimum mass solar nebula
(Hayashi 1981), the local surface density and sound ve-
locity of the PPD are ε×143 g cm−2 and 4.58×104cm s−1,
respectively, where ε is a scaling factor representing the
reduction of the surface density due to gas dissipation.
With these values, we obtain the mass infall rate as
f =(1.3×10−3ε(cid:16) r
RJ(cid:17)−1
0
g cm−2 s−1
(r ≤ 20RJ)
(r > 20RJ) ,
(2)
where RJ is one Jupiter radius. As the power-law index
of the mass infall rate drops outside ∼ 0.04 Hill radii
from the planet (Tanigawa et al. 2012), we set f = 0
at r > 20RJ. The initial value ε = 1 corresponds to the
beginning of the mass infall and a smaller value indicates
a smaller mass infall rate. Since the viscous timescale of
the CPD is sufficiently smaller than that of the PPD, we
treat ε as a constant here.
We adopt the following simplified form used in pre-
vious studies (Cannizzo 1993; Armitage et al. 2001) to
estimate the temperature structure, that is
∂Tc
∂t
=
2 (Q+ − Q−)
cpΣ
∂Tc
∂r
,
− vr
(3)
where Q+ = (9/8)νΣΩ2
K represents the viscous heat-
ing, Q− = σ (1 + 3/8κΣ) T 4
is the radiative cooling,
c
σ is the Stefan-Boltzmann constant, and vr is the ra-
dial velocity. Further to this, the opacity κ is given by
Bell & Lin (1994), summarized in Table 1(see also table 1
of Kimura & Tsuribe 2012) as
κ = κ0 ρa T b ,
(4)
where ρ is the density and T is the temperature. The
specific heat is given by cp ≃ 2.7 R/¯µ, where R is the
ideal gas constant and ¯µ = 2.34 is the mean molecular
weight.
Moon Migration in Weakly Accreting CPDs
3
opasity regime
Ices
Sublimation of ices
Dust
Sublimation of dust
Molecules
κ0
(cm2 g−1)
2 × 10−4
2 × 1016
1 × 10−1
2 × 1081
1 × 10−8
a
0
0
0
1
2/3
b
2
-7
1/2
-24
3
from (K)
Temperature range
to (K)
166.8
202.6
0
166.8
202.6
2286.7ρ2/49
2029.7ρ1/81
2286.7ρ2/49
2029.7ρ1/81
10000.0ρ1/21
Bell and Lin opacity from Table 1 of Kimura & Tsuribe (2012)
Table 1
t
i
r
c
,
e
x
10-4
10-5
10-6
10-7
10-8
10-9
1
T=100K
T=1000K
T=10000K
10
r [RJ]
100
Figure 1. Critical ionization degree needed to sustain the MRI
at the midplane at T = 100 K, 1000 K, and 10,000 K for β0 = 105.
We solve Eqns. (1) and (3) numerically with bound-
ary conditions of zero torque and vanishing temperature
gradient at the inner and outer boundaries. The tem-
perature of the PPD at the location of the sub-disk is
assumed to be T = 123 K. We set the temperature of the
CPD to this value whenever the calculated temperature
is lower than this.
2.2. Origin of Viscosity
In this section, we explain how to obtain an esti-
mate for the kinematic viscosity coefficient. The best-
known origin of effective viscosity in accretion disks is
via MRI turbulence. There are two criteria for the
MRI that must be fulfilled for the instability to be ac-
tive: the disk gas must be ionized enough to be coupled
with magnetic field, and the magnetic field is not too
strong (Balbus & Hawley 1991; Sano & Miyama 1999;
Okuzumi & Ormel 2013). We consider the MRI if the
following two conditions are satisfied:
1. the Elsasser number, Λ ≡ v2
vAz ≡ Bz/√4πρ being the vertical component of
the Alfv´en velocity, and where η is the magnetic
diffusivity, and
Az/ηΩK > 1, with
2. the z component of plasma beta defined in terms
of net magnetic flux, βz = 2c2
Az & 2000; see
Okuzumi & Ormel (2013) and Fujii et al. (2014)
for details.
s /v2
As suggested by the results of Fujii et al. (2011, 2014),
Turner et al. (2014), and Keith & Wardle (2014), sub-
disks are not likely to widely sustain well-developed MRI
turbulence in the absence of thermal ionization.
If the disk is not subject to sustaining MRI turbulence,
material may pile up and the disk eventually becomes
gravitationally unstable. Accordingly, if the Toomre pa-
rameter,
Q ≡
ΩKcs
πG Σ
,
(5)
(with G the gravitational constant) becomes smaller
than a value of 2, we employ an effective viscosity of
αGI ≡ exp(Q−4) (Zhu et al. 2010; Takahashi et al. 2013).
In such a case, the disk can be easily heated up provided
that gravitational energy is converted into heat.
It is
widely assumed that thermal ionization can trigger the
MRI if the temperature exceeds about 1000 K; however,
in CPDs this is not always the case. First of all, the
ionization fraction obtained from the Saha equation de-
pends on density and a gravitationally unstable disk nat-
urally has a high surface density. Secondly, for a CPD to
sustain the MRI, the required ionization fraction is com-
paratively higher than that of a PPD even if the density
is the same.
As is mentioned in Fujii et al. (2014), this is because
typical length scale is orders of magnitude smaller in a
sub-disk. Thus, the critical temperature is higher than
1000 K. If we adopt η = 234(T /1K)1/2 x−1
cm2s−1 where
xe ≡ ne/nn is ionization degree and ne and nn are re-
spectively number density of electron and neutral gas
(Blaes & Balbus 1994), the condition for sufficient ion-
ization to sustain the MRI at midplane can be given as
e
Λ =
2c2
s xe
234pT /1K ΩKβ0
> 1,
(6)
where β0 is the plasma beta at the midplane. From
this equation, we can derive the
ioniza-
tion degree needed to be MRI-active as xe,crit =
critical
s ).
234pT /1KΩKβ0/(2c2
As an illustration of this, Fig. 1 shows the critical ion-
ization degree needed to have the MRI at the midplane
for a disk with β0 = 105. The plot is made for the
disk temperatures of 100 K, 1000 K, and 10,000 K, respec-
tively. Note that, with the "gas-starved" case of surface
density as the lower limit, Σ = 100 (r/20RJ)−3/4 g cm−2
(Canup & Ward 2002; Sasaki et al. 2010), midplane ion-
ization degree at 10 RJ, when T = 1000 K is assumed
is only about 10−12. One can easily see from Fig. 1
that this value is far below the critical value, xcrit, at
the respective radius. The temperature needed to obtain
xe > xe,crit for this case is about 2000 K. Obviously, if the
surface density is larger, higher temperature is required.
ionization
In the high-
of potassium, sodium, and magnesium.
temperature regime, where most of the metals are al-
ready ionized, atomic hydrogen also becomes a dominant
source of free electrons. At such temperatures, hydrogen
gas is already dissociated, so we employ the following
Saha equation for atomic hydrogen as well as for metals
In this paper, we consider the thermal
4
Fujii, Kobayashi, Takahashi & Gressel
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
2.2e12s
5.5e11s
8.6e9s
2.2e9s
5.5e8s
1.4e8s
4.4e7s
8.5e6s
10000
]
K
[
T
1000
2.2e12s
5.5e11s
8.6e9s
2.2e9s
5.5e8s
1.4e8s
4.4e7s
8.5e6s
10
r [RJ]
100
1
10
r [RJ]
Figure 2. Time evolution of surface density (left) and temperature structure (right) for the case with ε = 10−2 and αfloor = 10−4. Each
color shows the different time of the evolution. The uppermost line in each plot shows the steady-state value except within 2.2 RJ.
to solve for the ionization degree from collisions1:
xexK+
xK
=
xexNa+
xNa
=
xexMg+
xMg
=
xexH+
xH
=
2
h2
h2
(cid:19)3/2
nn (cid:18) 2πmekBTc
× exp(−5.0 × 104/Tc)
(cid:19)3/2
nn (cid:18) 2πmekBTc
2
× exp(−6.0 × 104/Tc)
(cid:19)3/2
nn (cid:18) 2πmekBTc
2
× exp(−8.9 × 104/Tc)
(cid:19)3/2
nn (cid:18) 2πmekBTc
2
× exp(−1.6 × 105/Tc)
h2
h2
(7)
(8)
(9)
(10)
where, xK etc. represent the number densities of each
species, and me, h, and kB are respectively electron mass,
the Planck constant, and the Boltzmann constant. With
thoes equations and charge neutrality, we finally obtain
the ionization degree as
1
Tc
h2
+ xNa
2 (cid:18) 2πmekBTc
(cid:19)3/4
xe =(cid:16) 2
nn(cid:17)
xK exp(cid:18)−5.0 × 104
(cid:19)
×h
exp(cid:18)−6.0 × 104
(cid:19)
+ xMg exp(cid:18)−8.9 × 104
(cid:19)
+ xH exp(cid:18)−1.6 × 105
(cid:19)i
Tc
Tc
Tc
1
2
.
(11)
We adopt Solar abundance multiplied by depletion factor
δ for metal species, xK = 9.87 × 10−8 δ, xNa = 1.60 ×
10−6 δ, and xMg = 3.67 × 10−5 δ, in our calculation.
1 These expressions are approximately correct when the ioniza-
tion fraction of each species is small, which is not appropriate for
K and Na in this case. Since their abundance is small, however,
the resulting ionization degree is not strongly affected.
If Λ > 1 is satisfied at the midplane, we set the vis-
cous parameter due to the MRI turbulence as αMRI =
1040/β0 + 0.015 (Okuzumi & Hirose 2011). Thus, we de-
notethe the Elsasser number at the midplane as Λ here
after. What if the ionization degree is not high enough to
sustain the MRI and the surface density is smaller than
the critical value to be gravitationally unstable? There
is always molecular viscosity, but it is negligibly small.
Gravitational interaction between the star, planet and
gas of a CPD can be an origin of angular momentum
transport (Rivier et al. 2012). Kelvin-Helmholtz-like in-
stabilities between sedimenting dust layers and gas can
generate turbulence which can be roughly estimated as
∼ 10−4 − 10−3. Moreover, it has been found that in
disks with imposed radial temperature gradients, the re-
sulting vertical shear can be a robust source of turbulence
via an analog of the Goldreich-Schubert-Fricke instability
(Nelson et al. 2013). Rigorously establishing the pres-
ence of the vertical-shear instability in CPDs will, how-
ever, require us to derive constraints on radiative cool-
ing timescales τcrit similar to the work by Lin & Youdin
(2015), who (in the context of PPDs) find the corre-
sponding criterion to scale with the disk thickness –
which is favorably large for the comparatively puffed-up
CPDs, implying less-restrictive conditions on τcrit.
There may be other ways of transporting angular mo-
mentum, however, those mechanisms contain uncertain-
ties and the specific value is not yet obtained. Thus, we
treat them via setting a floor value, αfloor, in this work.
In summary, we define the viscous parameter as
α = αMRI + αGI + αfloor
+ 0.015
β0
0
αMRI =(cid:26) 1040
αGI =(cid:26)exp(Q−4)
0
(if Λ > 1)
(if Λ ≤ 1)
(if Q < 2)
(if Q ≥ 2)
(12)
(13)
(14)
We take ε and αfloor as parameters and obtain structures
of CPDs based on Section 2.1 above.
3. RESULTING DISK STRUCTURE
Since the timescale for ε to drop is uncertain, we sim-
ply develop a disk for each parameter set from scratch
until it reaches a quasi-stationary state. First, we cal-
culate assuming only 1% of metals are in gas phase, i.e.
Moon Migration in Weakly Accreting CPDs
5
δ = 0.01. Figure 2 is an example of the formation of a
CPD with ε = 10−2 and αfloor = 10−4. Both surface
density and temperature increase with time. The upper-
most lines in Figure 2 show the values in the steady state
except for within ∼ 2RJ, where the radial profiles of the
surface density and temperature remain non-steady and
wiggle about.
In Fig. 3, the opacity, Elsasser number, Λ, and Q value
at the final stage in Fig. 2 are shown. In the quiescent
outer disk, the Elsasser number and Q parameter re-
main in the stable regime, that is Λ < 1 and Q > 2,
and accordingly α is determined by αfloor. The Elsasser
number occasionally exceeds unity in the inner disk and
that prevents the system from settling into a stationary
solution. Even if the inner disk remains time-variable,
the outer disk achieves a steady state independent of the
inner region. In the quiescent outer disk, a bump in the
surface-density profile forms because of the increase of
opacity due to the transition of the origin from dust sub-
limation to molecules. The dips in surface density (and
temperature) at the inner domain border are related to
the boundary conditions. Thus, we do not consider them
as a bump.
Figure 4 shows the surface density profiles (left col-
umn) and temperature (right column) of various disk
models once the outer disk has reached a steady state.
One can see that models with larger ε and/or smaller
αfloor generally become more massive and hotter. The
top panels of Fig. 4 show the surface density and tem-
perature structure for ε = 10−1. Because the ionization
degree reaches a near-critical value at the inner disk radii,
the gas accretion rate fluctuates in time (illustrated by
the shaded area in Fig. 4), and the disk structure is not
fully stationary in this regime. When MRI enabled by
thermal ionization is developed, depending on the set-
tings, the disk either ends up with a steady state with
smaller surface density or enters the gravito-magneto
limit cycle studied by Lubow & Martin (2012, 2013).
The middle panels of Fig. 4 show the disk structure for
ε = 10−2. For the case of αfloor = 10−3 (green dotted
line), relatively effective gas accretion keeps the surface
density smaller than the cases for αfloor = 10−4 (blue
solid line) and αfloor = 10−5 (pink dashed line). Since the
temperature does not become high enough with αfloor =
10−3, the MRI is not triggered by thermal ionization and
the whole disk settles into a steady state. Compared to
the top panels, values are generally slightly smaller.
The bottom panel of Fig. 4 illustrates a case where the
reduction factor decreases down to 10−3. Because the
infall flux is already small enough, the surface density
does not become that massive, and therefore the tem-
perature cannot be as high as supplying sufficient ion-
ization to sustain the MRI. We remark that the surface-
density range of our models is similar to the extended
outer disk of Mosqueira & Estrada (2003a), but temper-
ature is much higher in our models. For some parame-
ter sets, a bump can be seen in surface density that is
formed due to the change in opacity. A radial pressure
bump cannot be seen in a disk with small ε and/or large
αfloor because the surface density does not pile-up suf-
ficiently for the inner disk to transition into the higher
temperature regime required for the opacity transition.
Next, for the case when the temperature at which the
101
100
10-1
]
g
/
2
m
c
[
κ
Λ
10-2
1
100
10-1
10-2
10-3
10-4
10-5
1
105
Q
104
103
102
1
10
r [RJ]
10
r [RJ]
10
r [RJ]
Figure 3. Radial profiles of opacity (top), Elsasser number (mid-
dle), and Toomre's Q parameter (bottom panel) for the fiducial
case with ε = 10−2 and αfloor = 10−4.
thermal ionization plays a role for gas accretion is as
high as the one for grains to evaporate, we calculate disk
structures with δ = 1 corresponding to solar abundance.
We show the respective results in Fig. 5. For models with
ε = 10−3 (not show in the figure), we obtained identical
profiles as in the case with δ = 0.01 for all values of
αfloor. This is also true for the model with ε = 10−2 and
αfloor = 10−3. For all other models, the range in which
we do not obtain steady-state solutions slightly increase
because thermal ionization can provide more electrons at
the same temperature.
We conclude that the transition of the opacity regime
from sublimation of dust to molecules can produce in-
teresting structures in embedded CPDs for a variety of
reasonable disk models. In the following section, we dis-
cuss the orbital evolution of (proto-)satellites in the disk
models derived here.
4. CAPTURE OF SATELLITES IN RESONANT ORBITS
In the context of protoplanetary disks, disk-planet in-
teraction has been studied extensively in the literature.
6
Fujii, Kobayashi, Takahashi & Gressel
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
10
r [RJ]
10
r [RJ]
10
r [RJ]
1000
]
K
[
T
100
1
1000
]
K
[
T
100
1
1000
]
K
[
T
100
1
10
r [RJ]
10
r [RJ]
10
r [RJ]
Figure 4. Steady-state surface density (left) and temperature (right) for ε = 10−1, 10−2, and 10−3 (from top to bottom) with δ = 0.01
and various αfloor: dashed (pink), thick-solid (blue), dotted (green) lines show the case for αfloor = 10−5, 10−4, and 10−3, respectively.
Shaded regions are not time-steady and display jitter. As a reference, a typical gas-starved disk profile is also plotted (solid black line).
Planets are believed to migrate in the hosting disk by
exchanging their angular momentum with the disk. The
idea is also introduced in satellite formation (for in-
stance in Canup & Ward 2002, 2006; Sasaki et al. 2010;
Ogihara & Ida 2012). As for PPDs, the migration direc-
tion and speed depend on the satellite mass and the disk
structure, and the timescale is given as
tm =
1
b(cid:18) Ms
Mp(cid:19)−1(cid:18) Σr2
Mp(cid:19)−1(cid:18) cs
vK(cid:19)2
Ω−1
K ,
(15)
where Ms and Mp are the mass of the satellite and planet,
respectively, and b is a constant that determines the di-
rection and speed of the migration (Paardekooper et al.
2011; Kretke & Lin 2012; Ogihara et al. 2015).
In this
work, we only consider moons with circular orbits, and
we use the formula of Paardekooper et al. (2011)2 for the
2 In their formula, which is derived as a fit to a set of simulations
with a single perturbing body, the resulting torque from the entire
disk is considered. For systems of multiple embedded bodies, the
formula remains valid as long as all masses remain low enough,
such that non-linear wake interaction can be ignored.
migration constant b (see also Ogihara et al. 2015):
b =
2
γ {−2.5 − 1.7q + 0.1p
+1.1F (Pν)G(Pν )(cid:18) 3
+0.7 (1 − K(Pν))(cid:18) 3
2 − p(cid:19)
2 − p(cid:19)
F (Pν )F (Pχ)qG(Pν )G(Pχ)
+7.9
q − (γ − 1)p
γ
1.4
+(cid:18)2.2 −
γ (cid:19) [q − (γ − 1)p]
×q(1 − K(Pν ))(1 − K(Pχ))(cid:27) ,
(16)
where γ is the adiabatic constant, p ≡ −d ln Σ/d ln r,
and q ≡ −d ln T /d ln r. The expressions for F (P ), G(P ),
and K(P ) are furthermore given as
1.3(cid:19)2
F (P ) =(cid:16)1 +(cid:18) P
(cid:17)−1
(17)
Moon Migration in Weakly Accreting CPDs
7
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
105
104
]
2
m
c
/
g
[
Σ
103
102
101
1
10
r [RJ]
10
r [RJ]
1000
]
K
[
T
100
1
1000
]
K
[
T
100
1
10
r [RJ]
10
r [RJ]
Figure 5. Same as Fig. 4 but for δ = 1, that is, undepleted composition. The top panels are for ε = 10−1 and the bottom panels are for
ε = 10−2. The results for ε = 10−3 are not shown because they are identical to Figure 4.
ε
αfloor
Models 1 & 1'
Models 2 & 2'
Models 3 & 3'
Models 4 & 4'
Models 5 & 5'
10−1
10−1
10−2
10−2
10−3
10−4
10−3
10−5
10−4
10−5
Model parameters for the cases considered for orbital migration
Table 2
We assume γ = 1.4 and that the thermal diffusivity is the
same as ν (i.e., Pr=1). Thus, with the dimensionless half-
width of the horseshoe region, χs = 1.1/γ1/4pMsr/Mph
(where h is the scale-height of the disk),
Pν =
2
3r ΩKr2χ3
2πν
s
=
2
3
Pχ.
(20)
If b is negative, the satellite migrates toward the
planet and positive b means outward migration. We se-
lected disk models with a discernible bump in surface
density structure, as summarized in Table 2. In the fol-
lowing, we refer to the cases with δ = 0.01 as Models
1-5, and δ = 1 as Models 1'-5'. The radial distribution
of b for Models 1-5 with a satellite of Io mass is given in
Figure 6. If a Io-sized satellite that formed at outer radii
migrates inward, the migration halts at locations marked
with crosses in Fig. 6. Moons migrate all the way to the
planet in Model 5 because there is no location where the
G(P ) =
K(P ) =
P 3/2
16
8 (cid:1)3/4
25(cid:0) 45π
45π(cid:1)4/3
1 − 9
25(cid:0) 8
28 (cid:1)3/4
25(cid:0) 45π
45π(cid:1)4/3
25(cid:0) 28
1 − 9
16
P 3/2
45π
45π
P <q 8
P −8/3 P ≥q 8
P <q 28
P −8/3 P ≥q 28
45π
45π .
(18)
(19)
10
5
b
0
-5
-10
1
Model 1
Model 2
Model 3
Model 4
Model 5
10
r [RJ]
Figure 6. Migration coefficient for an Io-sized moon in the disk
model given in Section 2 for Models 1-5. The positions where the
migration stops for each model are indicated by crosses.
10
5
b
0
-5
-10
1
Model 1'
Model 2'
Model 3'
Model 4'
Model 5'
10
r [RJ]
Figure 7. Migration coefficient for an Io-sized moon for Models 1'-
5'. For the comparison, the positions where the first moons stop
migration for Models 1, 2, and 4 are marked with cross signs.
value of b changes from positive to negative as r increases.
8
Fujii, Kobayashi, Takahashi & Gressel
Figure 7 shows plots of b for Models 1'-5' with δ = 1,
that is, solar abundance. The location for the termina-
tion of moon migration is slightly further inside in Mod-
els 1', 2', and 4' compared to Models 1, 2, and 4, re-
spectively, whereas the convergence region disappears in
Model 3' as compared to Model 3. If a satellite migrates
and stays at b ≃ 0, the second satellite migrating from
the outer disk approaches the first one that is trapped
inside the location of convergent migration. If the mi-
gration timescale of the second satellite is longer than
the critical time scale, tcrit
m , at the location of the first
satellite, the second satellite is captured in a mean mo-
tion resonance. As mentioned in Ogihara & Kobayashi
(2013), the capture probability for higher-order reso-
nances is very low, and moreover the 2:1 MMR is the
outermost among first-order resonances – we hence ex-
clusively focus on this case. The critical time scale for
capture into 2:1 MMR of equal-mass satellites is given
by (Ogihara & Kobayashi 2013)
Migration timescale
Critical timescale
Model 1
Model 2
Model 3
Model 4
tm(14RJ) = 1500yr > tcrit
tm(21RJ) = 2600yr > tcrit
tm(7.6RJ) = 5000yr > tcrit
tm(12RJ) = 7600yr > tcrit
tm(9.8RJ) = 760yr > tcrit
tm(16RJ) = 1300yr > tcrit
tm(5.4RJ) = 2500yr > tcrit
tm(8.6RJ) = 4000yr > tcrit
m (8.5RJ) = 97yr
m (14RJ) = 200yr
m (4.8RJ) = 41yr
m (7.6RJ) = 82yr
m (6.2RJ) = 60yr
m (9.8RJ) = 120yr
m (3.4RJ) = 24yr
m (5.4RJ) = 49yr
Comparison between type-I satellite migration timescales, tm,
and critical timescales, tcrit
m , for capture into MMR.
Table 3
Model 1
Model 1'
Model 2
Model 2'
Model 3
Model 4
Model 4'
tcrit
m = 2.5 × 104(cid:18) Ms
MIo(cid:19)−4/3(cid:18) Mp
MJ(cid:19)4/3
TK ,
(21)
Galilean moons
where MIo is the mass of Io and TK is the orbital period of
the satellite. Here, we only consider satellites with equal
masses because Galilean satellites have similar masses.
In Model 1, the first satellite is located at about 8.5RJ
after the termination of migration. A satellite in 2:1 reso-
nance with the satellite at 8.5RJ has an orbit at approx-
imately 14RJ. The corresponding migration timescale
is tm(14RJ) ≃ 1500 yr, which is longer than the critical
time scale for the first satellite, tcrit
m (8.5RJ) = 97 yr. This
means that the second satellite is captured in the reso-
nance. Similarly, the third satellite can be captured in
the 2:1 resonance of the second satellite at 21RJ because
tm(21RJ) ≃ 2600 yr > tcrit
m (14RJ) ≃ 200 yr. In this way,
we successfully build-up a system in Laplace resonances.
The positions where the first satellite terminate for
Models 2-4 are 4.8RJ, 6.2RJ, and 3.4RJ, and the 2:1
resonance orbits of these are 7.6RJ, 9.8RJ, and 5.4RJ,
respectively. The orbits in 2:1 resonance with the sec-
ond satellites are 12RJ, 16RJ, and 8.6RJ, respectively.
As summarized in Table 3, the migration timescale of
each of these orbits is larger than the critical timescale
for capture in the mean motion resonance. Thus, we can
also obtain systems in the Laplace resonance with Mod-
els 2-4, as well as Model 1. Similarly, we can form those
systems in Models 1', 2' and 4'. The comparison of orbits
of Models 1-4, 1', 2' and 4' with the Galilean moons are
given in Figure 8. Note that the orbits of the resonant
three moons are located on the same slope of the surface
density profile in all models. One can see that Model 3
has a similar set of orbits with the inner three moons of
the Galilean system.
5
10
15
20
r [RJ]
Figure 8. Comparison of the resonant orbits of the satellites ob-
tained in our models with those of the Galilean moons.
Galilean Moons
Model 3
Mass (1025 g) Orbit (RJ) Orbit (RJ)
Io
Europa
Ganymede
8.93
4.80
14.8
5.9
9.4
15.0
6.2
9.8
16
Table 4
Summary of satellite properties
this resonance, the orbits are locked and the moons mi-
grate together as a system; the separations of the bodies
adjust accordingly, when the whole system moves radi-
ally during the evolution of the CPD.
]
1
-
s
[
Ω
10-3
10-4
10-5
10-6
10-7
1
10
r [RJ]
5. DISCUSSION
Figure 9. Angular velocity of the disk for Model 4. Keplerian
frequency is also plotted in the dotted line.
We summarize the orbits of satellites in our Model 3
along with the mass and orbits of the inner three Galilean
moons that are in the Laplace resonance in Table 4.
Systems in other models are more compact or spread-
out compared with the Galilean moons, but most impor-
tantly, the moons are in the 4:2:1 MMR. Once they are in
Figures 4 and 5 show that the disk is quite hot at
this stage. At such high disk temperatures, the ra-
dial pressure gradient may lead to sub-Keplerian rota-
tion velocities. Actually, as Fig. 9 shows, the angu-
lar velocity is smaller than the Keplerian value in the
Moon Migration in Weakly Accreting CPDs
9
outer part of the disk. The angular velocity is calcu-
lated as Ω = ΩK(1 − η)1/2, where η = −(rΩ2
Kρ)−1∂P/∂r
(Takeuchi & Lin 2002). We assumed Keplerian rotation
profiles when we derived disk models. However, since
Equation (1) is only sensitive to the radial slope of the
angular velocity, the assumption is expected to be ac-
ceptable.
Hot CPDs are suggested by Keith & Wardle (2014),
Zhu (2015), and Szul´agyi et al. (2016), but it may be
difficult to form icy satellites in such an environment.
Since the outer disk is cooler, moons may gain icy mate-
rials simply by migrating in from larger radii. Although
Fig. 6 shows b is positive in r & 25RJ for Model 4, for
instance, bodies about ten times smaller than Io can mi-
grate all the way from the outer radii because b for them
remains negative at all radii. Another possibility is that
ice-rich planetesimals are captured when they enter into
the CPD. Tanigawa et al. (2014) found that the orbits of
sub-Io sized planetesimals captured in a CPD are highly
eccentric. They also found that 10 m or larger planetes-
imals can be efficiently captured in a CPD, thus those
bodies may grow into the size of present moons.
One problem is, however, whether the system can sur-
vive over the long-term evolution of the CPD. As mass in-
fall decreases, the temperature of the disk also decreases.
When the disk structure that traps the innermost body
disappears, the satellite system will start to migrate to-
ward the planet. Moons can survive if the CPD is quickly
cleared before they are lost into the planet. A rough esti-
mate of the viscous timescale of the disk is r2/ν ∼ 100 yr,
which is shorter than the migration timescale of the satel-
lites. However, the actual timescale for the surface den-
sity to become small enough not to affect satellite mi-
gration is most likely to be much longer than this esti-
mate. Clearly, this depends on how the disk dissipates
and many other unknown factors.
In order to obtain
a better understanding of how the infall terminates, we
need to further study the evolution of PPDs including
both gap formation and gas dissipation. In this work, we
adopted a mass infall rate derived from isothermal hydro-
dynamic simulations, however, Gressel et al. (2013) sug-
gested that taking magnetic field and radiative cooling
into account leads to a different mass infall rate, which
opens a perspective for future work.
Sasaki et al. (2010) and Ogihara & Ida (2012) found
that the existence of an inner cavity in a CPD can pre-
vent a moon system from being lost onto the planet. Pro-
vided the planet rotates differentially (given that it ac-
cretes material with non-negligible angular momentum
this is not unreasonable to assume) and maintains a con-
vective or turbulent sub-surface flow, it can be expected
to harbor an efficient planetary dynamo.
In this case,
such an inner cavity may form due to magnetospheric
truncation of the sub-disk by the planet's dipole mag-
netic field.
In the context of PPDs, not only photoe-
vaporative but also magnetically driven disk wind have
been reported to contribute to the formation of so-called
"transition disks" (Suzuki et al. 2010) with reduced sur-
face density at small radii. To explore such currently
unknown effects in the context of embedded sub-disks,
the configuration of magnetic field at the very vicinity of
the planet must be studied.
We have modeled massive and comparatively hot
CPDs by solving the time evolution of surface density
with mass infall from the parental PPD. The mass infall
flux was determined based on the high-resolution numer-
ical simulation of Tanigawa et al. (2012), where we have
also considered the reduction of the flux caused by the
dissipation of the PPD at the location of the sub-disk.
The temperature profile of the CPD is derived by the
balance of viscous heating and radiative cooling, as well
as the radial advection. Since the strength of viscosity is
uncertain in the absence of MRI, we employed a parame-
ter to determine the minimum value of the viscosity. We
considered the MRI when the Elssasser number exceeds
unity due to thermal ionization. We furthermore moni-
tored Toomre's Q parameter in order to consider effective
viscosity when the value becomes lower than about two.
When the evolution is governed by αfloor, the system set-
tles into a steady state.
In many previous studies, the critical temperature for
the onset of the MRI is assumed to be at about 1000 K.
As shown in Figure 1, however, we found that this is
not the case for massive CPDs. This is because of the
two reasons: (i) the ionization degree needed to sustain
the MRI in a CPD is higher than that in a PPD, and
(ii) thermal ionization is less effective in higher density
regions.
In our models, MRI is turned on by thermal
ionization only around T∼ 2000 − 3000 K.
We found that opacity transitions change the radial de-
pendence of the temperature structure, and especially, a
transition near 2000 K makes a bump in surface density
distribution. We estimated whether a moon migrating
toward the central planet can be trapped at such a loca-
tion. In the case of some of the parameter settings that
are referred to as Models 1-4, 1', 2', and 4', the surface-
density and temperature gradients were sufficiently steep
to stop the migration of a moon. Moreover, we have ex-
amined the migration timescales of the second and third
moons migrating inward and compared them to the criti-
cal timescale to be captured in a 2:1 MMR with the inner
moon. In all of Models 1-4, 1', 2', and 4', we obtained
systems in 4:2:1 mean motion resonance that is known for
inner three bodies of the Galilean system. The satellite
system obtained in our disk models may or may not sur-
vive until the dissipation of the CPD. In order to find out
the long term evolution of these systems, further studies
on mass infall from PPDs and on the origin of angular
momentum transport in CPDs are needed.
We thank the anonymous referee for a careful re-
port. We acknowledge Shigeo S. Kimura, Pablo Ben´ıtez-
Llambay, Shigeru Ida, Masahiro Ogihara, and Kazuhiro
D. Kanagawa for fruitful discussions and Edwin L.
Turner for encouraging comments. HK was supported
by Grants-in-Aid for Scientific Research (No. 26287101)
from Ministry of Education, Culture, Sports, Science and
Technology (MEXT) of Japan and by Astrobiology Cen-
ter Project of the National Institute of Natural Science
(NINS) (Grant Number AB281018). OG has received
funding from the European Research Council (ERC) un-
der the European Union's Horizon 2020 research and in-
novation programme (grant agreement No 638596).
6. SUMMARY
REFERENCES
10
Fujii, Kobayashi, Takahashi & Gressel
Armitage, P. J., Livio, M., & Pringle, J. E. 2001, MNRAS, 324,
705
Ayliffe, B. A., & Bate, M. R. 2009, MNRAS, 393, 49
Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214
Bell, K. R., & Lin, D. N. C. 1994, ApJ, 427, 987
Blaes, O. M., & Balbus, S. A. 1994, ApJ, 421, 163
Cannizzo, J. K. 1993, ApJ, 419, 318
Canup, R. M., & Ward, W. R. 2002, AJ, 124, 3404
-. 2006, Nature, 441, 834
Charnoz, S., Salmon, J., & Crida, A. 2010, Nature, 465, 752
Crida, A., & Charnoz, S. 2012, Science, 338, 1196
Estrada, P. R., & Mosqueira, I. 2006, Icarus, 181, 486
Fujii, Y. I., Okuzumi, S., & Inutsuka, S. 2011, ApJ, 743, 53
Fujii, Y. I., Okuzumi, S., Tanigawa, T., & Inutsuka, S. 2014, ApJ,
785, 101
-. 2013, MNRAS, 432, 1616
Mosqueira, I., & Estrada, P. R. 2003a, Icarus, 163, 198
-. 2003b, Icarus, 163, 232
Nelson, R. P., Gressel, O., & Umurhan, O. M. 2013, MNRAS,
435, 2610
Ogihara, M., & Ida, S. 2012, ApJ, 753, 60
Ogihara, M., & Kobayashi, H. 2013, ApJ, 775, 34
Ogihara, M., Kobayashi, H., Inutsuka, S.-i., & Suzuki, T. K. 2015,
A&A, 579, A65
Okuzumi, S., & Hirose, S. 2011, ApJ, 742, 65
Okuzumi, S., & Ormel, C. W. 2013, ApJ, 771, 43
Paardekooper, S.-J., Baruteau, C., & Kley, W. 2011, MNRAS,
410, 293
Perez, S., Dunhill, A., Casassus, S., et al. 2015, ApJ, 811, L5
Rivier, G., Crida, A., Morbidelli, A., & Brouet, Y. 2012, A&A,
Gressel, O., Nelson, R. P., Turner, N. J., & Ziegler, U. 2013, ApJ,
548, A116
779, 59
Hayashi, C. 1981, Progress of Theoretical Physics Supplement,
70, 35
Hyodo, R., Charnoz, S., Genda, H., & Ohtsuki, K. 2016, ApJ,
828, L8
Hyodo, R., Charnoz, S., Ohtsuki, K., & Genda, H. 2017, Icarus,
282, 195
Hyodo, R., Ohtsuki, K., & Takeda, T. 2015, ApJ, 799, 40
Keith, S. L., & Wardle, M. 2014, MNRAS, 440, 89
Kimura, S. S., & Tsuribe, T. 2012, PASJ, 64, 116
Klahr, H., & Kley, W. 2006, A&A, 445, 747
Kretke, K. A., & Lin, D. N. C. 2012, ApJ, 755, 74
Lin, M.-K., & Youdin, A. N. 2015, ApJ, 811, 17
Lubow, S. H., & Martin, R. G. 2012, ApJ, 749, L37
-. 2013, MNRAS, 428, 2668
Lunine, J. I., & Stevenson, D. J. 1982, Icarus, 52, 14
Machida, M. N., Kokubo, E., Inutsuka, S., & Matsumoto, T.
2010, MNRAS, 405, 1227
Martin, R. G., & Lubow, S. H. 2011, ApJ, 740, L6
Sano, T., & Miyama, S. M. 1999, ApJ, 515, 776
Sasaki, T., Stewart, G. R., & Ida, S. 2010, ApJ, 714, 1052
Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337
Suzuki, T. K., Muto, T., & Inutsuka, S.-i. 2010, ApJ, 718, 1289
Szul´agyi, J., Masset, F., Lega, E., et al. 2016a, MNRAS, 460, 2853
Szul´agyi, J., Mayer, L., & Quinn, T. 2017, MNRAS, 464, 3158
Szul´agyi, J., Morbidelli, A., Crida, A., & Masset, F. 2014, ApJ,
782, 65
Takahashi, S. Z., Inutsuka, S.-i., & Machida, M. N. 2013, ApJ,
770, 71
Takeuchi, T., & Lin, D. N. C. 2002, ApJ, 581, 1344
Tanigawa, T., Maruta, A., & Machida, M. N. 2014, ApJ, 784, 109
Tanigawa, T., Ohtsuki, K., & Machida, M. N. 2012, ApJ, 747, 47
Tanigawa, T., & Watanabe, S.-i. 2002, ApJ, 580, 506
Turner, N. J., Lee, M. H., & Sano, T. 2014, ApJ, 783, 14
Zhu, Z. 2015, ApJ, 799, 16
Zhu, Z., Hartmann, L., & Gammie, C. 2010, ApJ, 713, 1143
|
1304.6184 | 1 | 1304 | 2013-04-23T07:20:50 | Layered convection as the origin of Saturn's luminosity anomaly | [
"astro-ph.EP",
"astro-ph.SR",
"physics.ao-ph",
"physics.flu-dyn"
] | As they keep cooling and contracting, Solar System giant planets radiate more energy than they receive from the Sun. Applying the first and second principles of thermodynamics, one can determine their cooling rate, luminosity, and temperature at a given age. Measurements of Saturn's infrared intrinsic luminosity, however, reveal that this planet is significantly brighter than predicted for its age. This excess luminosity is usually attributed to the immiscibility of helium in the hydrogen-rich envelope, leading to "rains" of helium-rich droplets. Existing evolution calculations, however, suggest that the energy released by this sedimentation process may not be sufficient to resolve the puzzle. Here, we demonstrate using planetary evolution models that the presence of layered convection in Saturn's interior, generated, like in some parts of Earth oceans, by the presence of a compositional gradient, significantly reduces its cooling. It can explain the planet's present luminosity for a wide range of configurations without invoking any additional source of energy. This suggests a revision of the conventional homogeneous adiabatic interior paradigm for giant planets, and questions our ability to assess their heavy element content. This reinforces the possibility for layered convection to help explaining the anomalously large observed radii of extrasolar giant planets. | astro-ph.EP | astro-ph |
Layered convection as the origin of
Saturn's luminosity anomaly
Jérémy Leconte1,2 & Gilles Chabrier2,3
1Laboratoire de Météorologie Dynamique, Institut Pierre Simon Laplace, Paris, France,
2Ecole Normale Supérieure de Lyon, CRAL (UMR CNRS 5574), 69364 Lyon, France,
3School of Physics, University of Exeter, Exeter, UK
E-mail: [email protected], [email protected]
As they keep cooling and contracting, Solar System giant planets radiate more en-
ergy than they receive from the Sun. Applying the first and second principles of
thermodynamics, one can determine their cooling rate, luminosity, and temperature
at a given age. Measurements of Saturn's infrared intrinsic luminosity, however, re-
veal that this planet is significantly brighter than predicted for its age1,2. This excess
luminosity is usually attributed to the immiscibility of helium in the hydrogen-rich
envelope, leading to "rains" of helium-rich droplets3−8. Existing evolution calcula-
tions, however, suggest that the energy released by this sedimentation process may
not be sufficient to resolve the puzzle9. Here, we demonstrate using planetary evolu-
tion models that the presence of layered convection in Saturn's interior, generated,
like in some parts of Earth oceans, by the presence of a compositional gradient, sig-
nificantly reduces its cooling. It can explain the planet's present luminosity for a
wide range of configurations without invoking any additional source of energy. This
suggests a revision of the conventional homogeneous adiabatic interior paradigm for
giant planets, and questions our ability to assess their heavy element content. This
reinforces the possibility for layered convection to help explaining the anomalously
large observed radii of extrasolar giant planets.
Many arguments suggest the existence of compositional gradients in giant planet interiors, as a
consequence of either their formation process or their cooling history10,11,12,13. However, the effect of
this gradient on the thermal evolution of the planet is usually neglected for sake of simplicity. When a
vertical gradient of heavy elements is present in a fluid, the resulting mean molecular weight gradient
(decreasing upward) can prevent large scale convection to develop by counteracting the destabiliz-
ing effect of the temperature gradient. The complex interaction between advection and diffusion of
1
heat and atomic concentration can trigger a hydrodynamical instability, called the double diffusive
instability14, leading to a regime of double diffusive convection (also called semi convection). This
process significantly affects heat and element transport, as observed in Earth oceans. The instability
leads to either a state of homogeneous double diffusive convection where diffusive transport is only
modestly enhanced by small scale turbulence, or to a state of "layered convection", with numerous,
small convective layers separated by thin, diffusive interfaces corresponding to discontinuities in the
composition of the fluid, in which transport is enhanced more significantly15,16.
Numerical calculations show that, for relevant thermal and atomic diffusivities, both scenarios are
possible, depending on the ratio of the compositional to superadiabatic temperature gradients17. In
planets, however, this latter quantity is not a free parameter but is imposed by the (measured) energy
flux to be transported in the planet. It can be shown that this flux is too high to be transported by
the diffusive regime of the double diffusive instability (said differently, the thermal gradient needed
to transport the internal flux over the whole planet would be too high to be stabilized by the heavy
element gradient13). Under planetary conditions, the system will thus most likely settle in the regime
of layered convection.
To study the impact of layered convection on giant planet thermal structure, we have developed
an analytical model of the heat transport in such a medium13. Extending the usual mixing length
formalism18 to layered convection, we can calculate the internal temperature gradient as a function
of the local properties of the fluid and of a unique free parameter, namely the characteristic size of
the convective layers, l, or equivalently the corresponding dimensionless mixing length parameter,
α = l/HP, where HP is the pressure scale height in the fluid (see Methods).
Figure 1: Evolution of the total internal energy (Etot ≡ Eg + Eth + Erot) with time for various Saturn
models. The solid curve corresponds to the reference adiabatic model. The long dashed, dashed and
dotted curves are models with layered convection with α = 10−2.5, 10−3 and 10−3.5, respectively.
Because of the reduced intrinsic luminosity caused by layered convection, these models cool more
slowly and keep the memory of their initial energetic state much longer.
2
10(cid:45)40.0010.010.1110(cid:45)2.5(cid:45)2.0(cid:45)1.5(cid:45)1.0AgeGyrTotalEnergyEtot1035JAdiabaticΑ(cid:61)10(cid:45)2.5Α(cid:61)10(cid:45)3Α(cid:61)10(cid:45)3.5A first confirmation of the viability of this scenario for giant planet interiors has been given by
the fact that structure models with layered convection matching all the observational mechanical
constraints of Jupiter and Saturn (radius, gravitational moments), as well as the atmospheric helium
and mean heavy element abundances, have been obtained for a rather wide range of mixing length
parameters, namely α ∈ [10−2 − 4× 10−6] for Saturn and α ∈ [10−2 − 3× 10−5] for Jupiter13. These
constraints suggest that layered convection is favored over homogeneous double diffusive convection
(Supplementary Information). However, the question remains whether layered convection can also
explain the present luminosity of the giant planets.
In order to answer this question, we have computed the thermal evolution of planets with layered
convection. Because departure from adiabaticity can be significant in these interiors, usual isen-
tropic evolution calculations2 cannot be used. Instead, the evolution is computed by integrating the
intrinsic luminosity, Lint = −dEtot/dt, where the total energy Etot = Eg +Eth +Erot includes the grav-
itational, thermal and rotational energies (see Methods). The size of the convective/diffusive layers
is assumed to have reached an equilibrium value13,15 and is thus kept constant. For all the models
(i.e. the reference, homogeneous/adiabatic case and the semi convective ones, for any given α), the
amount of heavy elements is kept constant and equal to the one that fulfills the gravitational moment
constraints13 (Supplementary Information).
Figure 2: Cooling sequences of Saturn models with layered convection. a) Effective (Teff; solid
curves) and intrinsic (Tint; dashed curves) temperature (see Methods) evolution in time for the adia-
batic reference model (black) and three models with layered convection (α = 10−2.5, 10−3 and 10−3.5
from dark to light red). b) Zoom on present era. Dots show the observed effective temperature. At
early ages the effective temperature of models with layered convection is lower due to inefficient con-
vection. After a few hundred million years, these models become brighter due to the release of the
excess of energy stored from the initial state.
As illustrated in Fig. 1, our results show that starting from the same total internal energy, the time
3
(cid:230)(cid:230)0.010.1110100200150AgeGyrEffectiveTemperatureK4.04.24.44.64.85.0859095100105110115AdiabaticΑ(cid:61)10(cid:45)2.5Α(cid:61)10(cid:45)3Α(cid:61)10(cid:45)3.5abrequired to release a given amount of this energy is significantly longer in the layered case than in the
homogeneous adiabatic one. This does not imply, however, that the intrinsic luminosity of the object
will necessarily be larger at any given time. On the contrary, at early ages, objects with a layered
convection zone are far less luminous because of the reduced heat flux release. However, after some
time, the decrease in luminosity imposed by layered convection is overpowered by the increase of
internal energy to be released and planets with layered convection eventually become more luminous
than the ones with adiabatic interiors (Fig. 2). In Saturn's case, this "crossing time" occurs before
Saturn's present age so that layered convection yields a larger luminosity at 4.6 Gyr. As shown
in Fig. 2 and Fig. 3, if the size of the convective layers is about 2-3×10−3 pressure scale heights,
this process properly leads to the currently observed effective temperature, radius, and gravitational
moments of the planet without any additional energy source such as helium rains. The radius evolution
of these models is shown in Supplementary Fig. 5. For this range of layer sizes, the present interior
temperature of Saturn is only modestly affected compared to the adiabatic case (see Supplementary
Fig. 6). The temperature during the early evolution, however, is much higher.
Figure 3: Impact of the size of the layered convection zone on Saturn's cooling sequence. a) Effective
(Teff; solid curves) and intrinsic (Tint; dashed curves) temperature evolution for the adiabatic model
(black curve) and three scenarios with layered convection. b) Zoom on the present era. From darker
to lighter red curves we have, i) our baseline scenario with layered convection present throughout the
gaseous envelope of the planet (MLC = 0.8M; α = 2× 10−3), ii) a scenario where only the inner 50%
of the envelope exhibits layered convection (MLC ≈ 0.5M; α = 2× 10−4), iii) a case with a layered
convection shell between 0.5 and 0.7M (MLC ≈ 0.2M; α = 10−4).
Fig. 2 suggests that Saturn models with convective/diffusive layers smaller than 2-3×10−3 pres-
sure scale heights will be too bright at the age of the Solar System. This is not necessarily true.
If layered convection develops only within a restricted fraction of the planet, models with a lower
convective efficiency (lower α) in the layered zone and with an efficient convection everywhere else
4
(cid:230)(cid:230)0.010.1110100200300150AgeGyrEffectiveTemperatureK4.04.24.44.64.85.0889092949698AdiabaticMLC(cid:61)0.5MMLC(cid:61)0.8MMLC(cid:61)0.2Mabcan also reproduce Saturn's proper cooling timescale. Our results are valid for very different sizes
of the layered convection zone. Without covering the whole parameter space, we illustrate this by
showing an evolution track for Saturn where layered convection is restricted to the inner 50% in mass
of the planet above the core, with α = 2× 10−4, and an other track where layered convection oc-
curs in a shell between 50% and 70% of the planet's mass (Fig. 3). This latter case shows that the
layered convection region does not need to extend down to the core and could be present around the
molecular/metallic transition region or near an immiscibility region in the gaseous envelope.
We have also tested the sensitivity of the model to the initial conditions. Indeed, scenarios of
giant planet formation by core accretion do not provide so far a clear prescription for these initial
conditions. Large uncertainties in various parameters (gas-to-dust ratio, opacities,...) and the lack of a
proper treatment of the radiative shock during the gas runaway accretion phase, for instance, hamper
a proper determination of the planet's initial energy content19,20. Whereas these (unknown) initial
conditions become inconsequential after about a few tens of million years for an adiabatic planet of
Saturn's mass20, it takes significantly longer (up to a few billion years) for a planet where convection
is inefficient, because of the lower luminosity and thus the longer thermal, so-called Kelvin-Helmholtz
timescale (τKH ∝ 1/Lint). Initial conditions thus have a stronger impact on the cooling timescale of
planets with layered convection.
As shown in Fig. 4, however, these different initial conditions do not qualitatively affect our re-
sults, as there is always a possible trade-off between the initial energy content of the nascent planet (at
the end of the runaway accretion shock) and the thickness or number of convective-diffusive layers.
As seen in the figure, evolution of models with layered convection can adequately reproduce Saturn's
presently observed luminosity with very low initial energy content provided sufficiently small convec-
tive/diffusive layers. Given the current uncertainties of formation models, thermal evolution of Saturn
with layered, inefficient convection can thus meet the observational constraints for a wide variety of
initial conditions. Interestingly enough, the same layered convection mechanism, which inhibits the
heat flow, also provides a plausible explanation to the observed low luminosity of Uranus11. In this
case, however, the "crossing time" discussed above should be larger than the age of the Solar System.
Because the present luminosity of Jupiter is smaller than the one that is predicted by homogeneous,
adiabatic evolution models2 (see Supplementary Fig. 7), observations do not seem to support the
presence of layered convection over the whole planet at the present epoch. This does not preclude,
however, the possibility for this process either to be present in part of the planet or to have occurred
in the past, but the planet's gaseous envelope to have been homogenized early enough, due to the
more vigorous convective flux. The flux emitted by Jupiter during its cooling history is large enough
to redistribute a significant fraction of a massive initial core (of up to 15− 20M⊕) within less than
4.6 Gyr21. Layered convection in this planet could then have stopped after the erosion of the whole
core, or at least of its soluble components22,23. This would be consistent with both i) the large initial
core mass needed to form the giant planet on the right timescale24 and ii) Jupiter's present smaller
core21,25,13. Jupiter and Saturn may thus have experienced quite different cooling histories and end
up with dissimilar interiors.
Layered convection does not preclude either the possibility of a phase separation between hydro-
gen and either volatiles or helium in Saturn's interior. In fact, a phase separation favors the occurrence
of layered convection26. Indeed, the strong compositional gradients (discontinuities) due to the pres-
5
Figure 4: Past luminosity of Saturn's model with layered convection. Evolutionary tracks of the
intrinsic luminosity (in units of the presently measured luminosity) of Saturn models integrated back-
ward in time from presently observed conditions with different mixing length parameters (α). All the
models with α > 2× 10−3 have too short cooling timescales and cannot be integrated backward for
4.6 Gyr. Even for quite low initial energy contents, i.e. a low initial intrinsic luminosity, models with
a sufficiently inefficient convection can reproduce the currently observed luminosity of Saturn.
ence of an immiscibility region provide very favorable conditions for the development of the double
diffusive instability and thus of layered convection. Our present treatment of layered convection does
not take into account the energetic contribution of a possible phase separation. In that case, however,
the energy release due to the phase separation does not need to be important, as the main role of
immiscibility would be to trigger the double diffusive instability. Indeed, as both layered convection
and phase separation increase the planet's cooling time, there is always a possible trade off between
the respective contributions of these two processes to the planet's cooling rate. Note, however, that
layered convection yields a hotter interior than the adiabatic evolution, favoring miscibility, especially
in the past (see Supplementary Fig. 6). As a smaller immiscibility region, if any, is predicted for he-
lium in Jupiter7,8, this scenario would also explain the dichotomy revealed here between Jupiter and
Saturn.
The cooling history of our Solar System giants, and of extrasolar giant planets in general, might
thus be more complex than assumed by the conventional, simplistic homogeneous, adiabatic interior
structure and evolution paradigm. The amazing diversity of the observed properties of extrasolar
planets - in particular the anomalously large observed radii of hot jupiters27 that cannot be fully
explained by the various proposed heating mechanisms28,29 - combined with the luminosity anomalies
of our own Solar System giants, clearly suggest a significant revision of this paradigm and point to a
broader, more complex picture of solar and extrasolar giant planet structure, composition and thermal
evolution, with a direct impact on giant planet formation conditions.
6
(cid:45)4(cid:45)3(cid:45)2(cid:45)10125102050100TimebeforepresentGyrIntrinsicLuminosityLSatΑ(cid:61)2.10(cid:45)3Α(cid:61)10(cid:45)3Α(cid:61)10(cid:45)3.5Α(cid:61)10(cid:45)4Methods
Layered convection and thermal gradient
To study the impact of layered double diffusive convection on the interior thermal structure, we have
developed an analytical model of the heat transport in such a medium13, which is briefly outlined
below.
The fluid is assumed to be composed of a large number of small convective layers of height l,
separated by thinner diffusive interfaces. The transport in the convective layers is described by a
parametrization similar to the standard mixing length theory of convection, for which the mixing
length is chosen to be equal to the height of the layers (l). The temperature gradient (∇T ≡ ∂ lnT
∂ lnP ),
hence the superadiabaticity in the gas, ∇T − ∇ad, is given as a function of the flux to be transported,
of the thermodynamic properties of the medium and of l.
In the diffusive interfaces, the thermal gradient is equal to the gradient needed to transport the
whole intrinsic flux by diffusive processes alone (∇d). The thermal size of these diffusive interfaces,
δT , can be estimated by the fact that the convective timescale, given by the inverse of the Brunt-
Väisälä frequency (τc = N−1
T /κT ,
where κT is the thermal diffusivity of the medium). Thus, δT =(cid:112)κT /NT , and the only remaining
T ) must be equal to the diffusive timescale in the interface (τd = δ 2
free parameter is the height of the convective layers, or equivalently the mixing length parameter,
α = l/HP, where HP is the pressure scale height in the fluid.
Finally, the mean temperature gradient (cid:104)∇T(cid:105) is a linear combination of the two aforementioned
temperature gradients weighted by the relative size of the convective and diffusive regions, and de-
pends on the size of the convective/diffusive cells13 (i.e. on α).
Planetary model sequences and time integration
Thermal evolution tracks of a planet of a given mass M, composition (symbolically denoted by an
array, X, encompassing all the information needed to know the abundances of all the materials at all
depths, including the core), angular momentum (J), and thermal structure (parametrized here by the
mixing length parameter α) yield relations of the type
(cid:0)T 4
(cid:1) = Lint(M, X,α,J,t),
int = 4πR2σSB
Lint ≡ 4πR2σSBT 4
eff − T 4
Etot ≡ Eth + Eg + Erot = Etot(M, X,α,J,t),
R = R(M, X,α,J,t),
eq
(1)
where Tint and Teff are the intrinsic and total effective temperatures, σSB the Stefan-Boltzmann con-
stant and R the radius of the planet. The mixing length parameter is assumed constant as convec-
tive/diffusive layers are assumed to have reached a state of equilibrium13,15. Without any further
guidance from tridimensional hydrodynamical simulations covering a large enough spatial and tem-
poral domain, it is sensible to start with the most simple assumption.
The equilibrium temperature (Teq) is the temperature that the planet would have if its intrinsic
7
luminosity were zero and represents the contribution to the solar flux through
σSBT 4
eq ≡ (1− A)L(cid:12)/(16πa2),
(2)
where L(cid:12) is the solar luminosity, a the orbital distance of the planet, and A its Bond albedo. Through-
out this study, the values used for the equilibrium temperature are Teq = 109K and 91K for Jupiter
and Saturn respectively. This corresponds to an absorbed luminosity of 5.14×1016 W (1.11×1016 W)
and a Bond albedo of 0.37 (0.32) for Jupiter (Saturn). Within the observational uncertainties, those
values are consistent with observations.
The relations described by (1) can be derived with a grid of atmosphere models. These grids of
atmospheric boundary conditions provide us with the temperature at the 10 bar level (or any reference
level deeper than the photosphere), T10(Tint,g), as a function of the intrinsic effective luminosity (Tint),
and of the gravity, g ≡ GM/R2. The details of the functional form used to represent the atmospheric
boundary conditions are detailed in the Supplementary Information.
Then, for any arbitrary value of the luminosity, T10 is known, and the T − P profile is integrated
inward by integrating the standard structure equations for a rotating body (Supplementary Informa-
tion). Models are computed on a grid of luminosities. For each luminosity, the gravitational, thermal
( u being the specific internal energy) and rotational energy,
(cid:90) Gm
¯r
(cid:90)
Eg ≡ −
dm, Eth ≡
udm, Erot ≡ 1
2
J2
I ,
(3)
hence the total energy, are computed (see Supplementary Fig. 8). Rotational energy is computed
assuming negligible angular momentum loss, and is found to represents only a small fraction of the
total energy (≈ 2%). As the ANEOS equation of state provides us with u (the specific internal energy)
for the volatiles, the thermal contribution of the core is included in our calculations.
Finally, we are left with a tabulated form of functions of the type Etot = Etot(M, X,α,J,Lint), and
time dependence can be retrieved by choosing an initial value for the luminosity or initial energy
content and by integrating
.
(4)
(cid:12)(cid:12)(cid:12)(cid:12)M, X,α,J
−Lint =
∂ Etot
∂ t
8
References and Notes
1. Pollack, J. et al. A calculation of Saturn's gravitational contraction history. Icarus. 30, 111-128
(1977).
2. Fortney, J. J., Ikoma, M. Nettelmann, N., Guillot, T. & Marley, M. S. Self-consistent Model
Atmospheres and the Cooling of the Solar System's Giant Planets. Astrophys. J. 729, 32
(2011).
3. Salpeter, E. E. On Convection and Gravitational Layering in Jupiter and in Stars of Low Mass.
Astrophys. J. 181, L83-L86 (1973).
4. Stevenson, D. J. & Salpeter, E. E. The dynamics and helium distribution in hydrogen-helium
fluid planets. Astrophysical Journal Supplement Series. 35, 239-261 (1977b).
5. Hubbard, W. B. & DeWitt, H. E. Statistical mechanics of light elements at high pressure. VII
- A perturbative free energy for arbitrary mixtures of H and He. Astrophys. J. 290, 388-393
(1985)
6. Pfaffenzeller, O., Hohl, D. & Ballone, P. Miscibility of Hydrogen and Helium under Astrophys-
ical Conditions. Phys. Rev. Lett. 74, 2599-2602 (1995).
7. Lorenzen, W., Holst, B. & Redmer, R. Demixing of Hydrogen and Helium at Megabar Pres-
sures. Phys. Rev. Lett. 102, 115701 (2009).
8. Morales, M. A. et al. Phase separation in hydrogen-helium mixtures at Mbar pressures. PNAS.
106, 1324 (2009).
9. Fortney, J. J. & Hubbard, W. B. Phase separation in giant planets: inhomogeneous evolution of
Saturn. Icarus. 35 (2003).
10. Stevenson, D. J. Cosmochemistry and structure of the giant planets and their satellites. Icarus.
62, 4-15 (1985).
11. Podolak, M., Hubbard, W. B. & Stevenson, D. J. Model of Uranus' interior and magnetic field.
In Uranus, University of Arizona Press, 29-61, (1991).
12. Chabrier, G. & Baraffe, I. Heat Transport in Giant (Exo)planets: A New Perspective. Astrophys.
J. Lett. 667, L81-L84 (2007).
13. Leconte, J. & Chabrier, G. A new vision of giant planet interiors: Impact of double diffusive
convection. Astron. & Astroph. 540, A20 (2012).
14. Stern, M. E. The salt-fountain and thermohaline convection. Tellus. 12, 172-+ (1960).
15. Radko, T. What determines the thickness of layers in a thermohaline staircase? J. Fluid Mech.
523, 79-98 (2005).
9
16. Rosenblum, E. P., Garaud, P., Traxler, A. & Stellmach, S. Turbulent Mixing and Layer Forma-
tion in Double-diffusive Convection: Three-dimensional Numerical Simulations and Theory.
Astrophys. J. 731,66-+ (2011).
17. Mirouh, G. M., Garaud, P., Stellmach, S., Traxler, A. L. & Wood, T. S. A New Model for Mixing
by Double-diffusive Convection (Semi-convection). I. The Conditions for Layer Formation.
Astrophys. J. 750, 61 (2012).
18. Hansen, C. J. & Kawaler, S. D. Stellar Interiors. Physical Principles, Structure, and Evolution.
Springer-Verlag Berlin Heidelberg New York (1994).
19. Pollack, J. B. et al. Formation of the Giant Planets by Concurrent Accretion of Solids and Gas.
Icarus. 124, 62 (1996).
20. Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P. & Lissauer, J. J. On the Luminosity
of Young Jupiters. Astrophys. J. 655, 541-549 (2007).
21. Guillot, T., Saumon, D. & Stevenson, D. J. The Interior of Jupiter. In Jupiter. The Planet,
Satellites and Magnetosphere.
22. Wilson, H. F. & Militzer, B. Solubility of Water Ice in Metallic Hydrogen: Consequences for
Core Erosion in Gas Giant Planets. Astrophys. J. 745, 54 (2012).
23. Wilson, H. F. & Militzer, B. Rocky Core Solubility in Jupiter and Giant Exoplanets. Phys. Rev.
Lett. 108, 1101 (2012)
24. Stevenson, D. J. Formation of the giant planets. Pla. Space Sci., 30, 755-764 (1982).
25. Saumon, D. & Guillot, T. Shock Compression of Deuterium and the Interiors of Jupiter and
Saturn. Astrophys. J. 609, 1170-1180 (2004)
26. Stevenson, D. J. & Salpeter, E. E. The phase diagram and transport properties for hydrogen-
helium fluid planets. Astrophys. J. Suppl. Ser. 35, 221-237 (1977a).
27. Miller, N. & Fortney, J. J. The Heavy-element Masses of Extrasolar Giant Planets, Revealed.
Astrophys. J. Lett. 736, L29 (2011)
28. Laughlin, G., Crismani & M., Adams, F. C. On the Anomalous Radii of the Transiting Extra-
solar Planets. Astrophys. J. Lett. 729, L7+ (2011)
29. Leconte, J., Chabrier, G., Baraffe & I., Levrard, B. Is tidal heating sufficient to explain bloated
exoplanets? Consistent calculations accounting for finite initial eccentricity. Astron. & Astroph.
516, A64+ (2010).
Correspondence and requests for materials should be addressed to J.L.
10
Acknowledgments
J.L. thanks J. Fortney for making his atmospheric grids available to us in electronic format. The
research leading to these results has received funding from the European Research Council under
the European Community's Seventh Framework Programme (FP7/2007-2013 Grant Agreement no.
247060)
Author contributions
J.L. carried out analytical calculations, developed the model and performed the numerical simulations.
G.C. suggested the idea and carried out analytical calculations. J.L and G.C. wrote the manuscript.
11
Supplementary Information
1 Layered convection or homogeneous double diffusive convec-
tion?
The transport efficiency derived from numerical simulations16,17 for a homogeneous double diffusive
medium (where transport is ensured by small scale turbulence) is equivalent, in our analytical model,
to a medium in a state of layered convection with α ≈ 10−8 − 10−9 (see Figure 3 of Leconte and
Chabrier13).
The fact that measured gravitational moments constrain α to be larger than 10−6 (13) thus seems
to favor the occurrence of layered convection over homogeneous double diffusive convection in Solar
System giant planets.
In addition, the evolutionary constraint on the luminosity discussed in the
present study, which constrain α to be larger larger than 10−4 (see Fig. 3), also seems to justify this
hypothesis.
2 Atmospheric model grids
The relations described by (1) can be derived with a grid of atmosphere models. These grids of
atmospheric boundary conditions provide us with the temperature at the 10 bar level (or any reference
level deeper than the photosphere), T10(Tint,g), as a function of the intrinsic flux exiting the planet
(parametrized by the intrinsic effective luminosity, Tint), and of the gravity, g ≡ GM/R2.
We use the recent atmospheric grids derived specifically for Jupiter and Saturn by Fortney et al.2.
However, as a planet with layered convection can be quite hot (high internal temperature) but with a
low luminosity (low effective temperature), our evolutionary tracks can start in the low g, low Tint part
of the parameter space that is usually not probed by adiabatic evolution models. The grids mentioned
above thus do not cover this range and extrapolation in g had to be used. The rather low dependence
of T10(Tint,g) on gravity justifies this procedure.
In order to retrieve the behavior of T10(Tint,g) in both the high and low Tint regime (arbitrarily
separated at Tmid), the numerical grids where fitted separately in the two regimes by functions of the
form
T10(Tint,g) = C + Kgβ (Tint − T0)γ .
(5)
A smooth function is recovered on the whole temperature range by interpolating linearly between the
two functions in the range [Tmid − ∆T,Tmid + ∆T ]. The 10 bar temperatures derived by this procedure
do not differ by more than 2-3% with respect to the tabulated values. Values used for the fitting
parameters are given in Table 1.
12
Table 1: Parameters used in Eq. (5) to fit the atmospheric grids of ref. 2.
Tmid (K)
∆T (K)
Jupiter
224
25
Saturn
225
25
T < Tmid
78.4
0.00550
-122.
-0.182
2.09
T > Tmid
-283.
198.
143.
-0.114
0.454
T < Tmid
62.7
0.00202
-97.1
-0.180
2.31
T > Tmid
-1680.
831.
141.
-0.0721
0.293
C
K
T0
β
γ
3 Planetary structure integration
When T10 is known, the T − P profile is integrated inward by integrating the standard structure equa-
tions for a rotating body
Gm
¯r4 +
∂ P
∂ m = − 1
4π
∂ ¯r
1
4π ¯r2ρ ,
∂ m =
∂ P
∂ T
T
∂ m =
∂ m
P
∇T ,
ω2
6π ¯r + ϕω (¯r),
(6)
(7)
(8)
where m is the mass enclosed in the equipotential of mean radius ¯r, ω the rotation rate of the planet,
ϕω (¯r) a second-order correction due to the centrifugal potential. As discussed above, ∇T is either
equal to ∇ad in the adiabatic case, or to (cid:104)∇T(cid:105) which depends on α in a layered convection zone. The
angular velocity is given by the fact that J ≡ Iω, I being the angular moment of inertia computed for
each model, is kept constant and equal to the present value.
A closure equation is provided by the equation of state (EOS) of the mixture along the planet's
interior profile. Such an EOS is generally given by the so-called ideal volume law for the mixture:
1
ρ =
X
ρX
+
Y
ρY
+
Z
ρZ
,
(9)
where X, Y and Z denote the mass fractions of H, He and heavy elements, respectively.
For each value of α, the core mass and the mass fraction of heavy elements as a function of
depth are considered constant in time and equal to the distribution that best matches the gravitational
moments and observational constraints13. For adiabatic, homogeneous models, the core masses and
uniform metal enrichment in the envelope are equal to ∼ 4M⊕ and Z = 0.11 in Jupiter and 25.5M⊕
and Z = 0.05 in Saturn. For models of Saturn with layered convection, the metal enrichments used
for a given mixing length parameter α are shown in 13 and the core mass used is ∼ 20.5M⊕. This
difference in the core mass between the homogeneous model and the layered ones is responsible for
13
the small energy difference remaining at low luminosity in Supplementary Fig. 8. Details about the
procedure and the equation of states used can be found in ref. 13 and reference therein.
Supplementary Fig. 7 shows adiabatic, homogeneous evolution tracks for Jupiter and Saturn com-
puted with this method. As expected, in the homogeneous case, our calculations are in good agree-
ment with previous results2: models cool to the observed effective temperature in ∼ 2.9 Gyr for
Saturn, and ∼ 5.2 Gyr for Jupiter. As we use the same atmospheric model grids, the slight differences
arising in the cooling times are probably due to the fact that our adiabatic models do not have ex-
actly the same core mass as the most recent calculations2, and that we take into account the thermal
contribution of this core.
14
Figure 5: Temporal evolution of the equatorial radius (in units of the measured present equatorial
radius of Saturn, 60 268 km) of an adiabatic Saturn model (solid black curve), and of three models
with layered convection (α = 3× 10−3, 10−3 and 3× 10−4 from dark to light red). The effect of the
deformation due to the fast rotation is taken into account. The time needed for the adiabatic model
fulfilling the constraints on the measured gravitational moments to cool down and contract to the
present radius is too short.
15
02468100.951.001.051.101.151.20AgeGyrReqRSatAdiabaticΑ(cid:61)3(cid:180)10(cid:45)3Α(cid:61)1(cid:180)10(cid:45)3Α(cid:61)3(cid:180)10(cid:45)4Figure 6:
Internal temperature-pressure profile of our baseline model with layered convection (red
curves; MLC ≈ 0.8M; α = 2× 10−3) at three different ages (dotted: 0.1 Gyr; dashed: 1 Gyr; long
dashed: 4.6 Gyr). The solid black curve represents the temperature profile inferred for the standard
adiabatic model. For comparison, the 1 Gyr profile of an adiabatic evolution is shown by the gray
dashed curve.
Figure 7: Effective (Teff; solid curves) and intrinsic (Tint; dashed curves) temperature as a function of
the planetary age for our adiabatic reference models of Jupiter (black) and Saturn (gray). Observed
effective temperatures are also shown as dots. As seen, our calculations in the conventional homo-
geneous adiabatic case are in good agreement with previous results2: models cool to the observed
effective temperature in ∼ 2.9 Gyr for Saturn, and ∼ 5.2 Gyr for Jupiter.
16
0.0010.010.1110100020005000100002000050000PressureMBarTemperatureK0.1Gyr1Gyr4.6Gyr(cid:230)(cid:230)(cid:224)(cid:224)01234566080100120140160AgeGyrTintKIntrinsic luminosity (Lint ≡ 4πR2 Fint) as a function of the total internal energy (Etot ≡
Figure 8:
Eg + Eth + Erot) for various sequences of Saturn models. The solid curve is the reference adiabatic
model. The long dashed, dashed and dotted curves are the α = 10−2.5, 10−3 and 10−3.5 sequences,
respectively. As expected, for a given internal energy content, the intrinsic luminosity of models
assuming layered convection decreases when α decreases and is significantly reduced compared with
the adiabatic case. This increases the Kelvin Helmholtz timescale, τKH ≡ GM2/(RLint), i.e. the time
needed for the object to loose the memory of its initial energy state.
17
(cid:45)2.0(cid:45)1.5(cid:45)1.0(cid:45)0.50.010171018101910201021TotalEnergyEtot1035JLuminosityLintWAdiabaticΑ(cid:61)10(cid:45)2.5Α(cid:61)10(cid:45)3Α(cid:61)10(cid:45)3.5 |
1104.4906 | 1 | 1104 | 2011-04-26T12:52:08 | Water Production by Comet 103P/Hartley 2 Observed with the SWAN Instrument on the SOHO Spacecraft | [
"astro-ph.EP"
] | Global water production rates were determined from the Lyman-{\alpha} emission of hydrogen around comet 103P/Hartley 2, observed with the SWAN (Solar Wind ANisotropies) all-sky camera on the SOHO spacecraft from September 14 through December 12, 2010. This time period included the November 4 flyby by the EPOXI spacecraft. Water production was 3 times lower than during the 1997 apparition also measured by SWAN. In 2010 it increased by a factor of ~2.5 within one day on September 30 with a similar corresponding drop between November 24 and 30. The total surface area of sublimating water within {\pm}20 days of perihelion was ~0.5 km^2, about half of the mean cross section of the nucleus. Outside this period it was ~0.2 km^2. The peak value was 90%, implying a significant water production by released nucleus icy fragments. | astro-ph.EP | astro-ph | Accepted for publication in Astrophysical Journal Letters
Water Production by Comet 103P/Hartley 2 Observed with the SWAN
Instrument on the SOHO Spacecraft
M.R. Combi1*, J.-L. Bertaux2, E. Quémerais2, S. Ferron3, J.T.T. Mäkinen4
1Dept. of Atmospheric, Oceanic and Space Sciences
University of Michigan
2455 Hayward Street
Ann Arbor, MI 48109-2143 USA
*Corresponding author: [email protected]
2LATMOS, CNRS/INSU,
Université de Versailles Saint-Quentin
11 Bd d'Alembert, 78280 Guyancourt FRANCE
3ACRI-st, Sophia-Antipolis, FRANCE
4Finnish Meteorological Institute, Box 503
SF-00101 Helsinki, FINLAND
M.R. Combi: [email protected]
J.L. Bertaux: [email protected]
E. Quémerais: [email protected]
S. Ferron: [email protected]
J.T.T. Mäkinen: [email protected]
ABSTRACT
Global water production rates were determined from the Lyman-α emission of hydrogen around
comet 103P/Hartley 2, observed with the SWAN (Solar Wind ANisotropies) all-sky camera on
the SOHO spacecraft from September 14 through December 12, 2010. This time period included
the November 4 flyby by the EPOXI spacecraft. Water production was 3 times lower than during
the 1997 apparition also measured by SWAN. In 2010 it increased by a factor of ~2.5 within one
day on September 30 with a similar corresponding drop between November 24 and 30. The total
surface area of sublimating water within ±20 days of perihelion was ~0.5 km2, about half of the
mean cross section of the nucleus. Outside this period it was ~0.2 km2. The peak value was
90%, implying a significant water production by released nucleus icy fragments.
2
1. INTRODUCTION
Comets are among the most volatile and least processed remnants of the nebula out of
which our solar system was formed 4.5 Gyr ago. The appearances of short-period, so-called
Jupiter-family comets (JFC) in the inner solar system are understood to originate from
gravitational perturbations of icy objects in the Kuiper belt outside the orbit of Neptune in a
process whereby they are passed down from Neptune to Uranus, then to Saturn, and finally to
Jupiter. From there many can be sent into orbits that have perihelia in the vicinity of 1 AU
(Fernandez 1980; Levison & Duncan 1997) and observable from the Earth. Typical JFCs have
periods in the range of 6 to 8 years and are then seen to be active from the Earth for a few
months around perihelion.
Comet 103P/Hartley 2 is one such JFC with an orbital period of 6.5 years and a
perihelion distance of 1.05 AU. It was discovered on June 4, 1986, by Malcolm Hartley at the
Siding Spring Observatory (Hartley 1986). It was seen during the 1991, 1997 and 2004
apparitions, though in 2004 it remained near superior conjunction on the other side of the sun
during most of its active perihelion phase. Observations from 1991 and 1997 indicated a
maximum water production rate of ~3 x 1028 molecules s-1 but a sharp drop with increasing
heliocentric distance (A'Hearn et al. 1995; Crovisier et al. 1999; Colangeli et al. 1999; Fink
2009; Combi et al. 2011). Infrared observations with the Spitzer Space Telescope (Lisse et al.
2009) made when the comet was not far past its most recent aphelion gave an estimate of its
mean radius to be rather small (0.57 km) implying the nucleus had to have a mostly active water
sublimating surface. 103P/Hartley 2 was ultimately chosen for a flyby during the extended
EPOXI mission for the Deep Impact spacecraft (A'Hearn et al. 2011), which had impacted and
made important measurements of JFC 9P/Tempel 1 in July 2005 (A'Hearn & Combi 2007).
3
Here we report the results of the analysis of observations of the hydrogen coma of comet
103P/Hartley 2 observed with the all-sky SWAN H Lyman-α camera on the SOlar and
Heliospheric Observatory (SOHO) spacecraft made during 3 months around the 2010 apparition
including the time period of the EPOXI flyby. From these we monitor the total global water
production rate of the comet as a function of time, which also provides important activity context
for other observations.
2. SOHO/SWAN OBSERVATIONS
The SOHO spacecraft has been in a halo orbit around the Earth-Sun L1 Lagrange point
observing the Sun and solar wind since its launch in late 1995. The Solar Wind ANisotropies
instrument is an all-sky scanning imager operating at the wavelength of neutral H Lyman-α at
121.6 nm (Bertaux et al. 1995). The main purpose of SWAN is to measure the Lyman-α
emission of the interstellar neutral hydrogen that streams through the solar system and is carved
out by the outflowing solar wind providing a global picture of anisotropies in the solar wind
flow.
Atomic hydrogen is the most abundant species in the atmosphere (or coma) of nearly all
comets. Most hydrogen is produced in a photodissociation chain originating with water
molecules and the OH radicals produced from water (Combi & Smyth 1988; Crovisier 1989).
Water is understood to be the most abundant volatile species in most comet nuclei and is
believed to control the activity of the coma when comets are within ~3 AU from the sun.
Because of their large hydrogen comae comets are easily observed by SWAN.
SWAN has observed the H Lyman-α coma of many comets (e.g., Bertaux et al., 1999,
Mäkinen et al. 2001, Combi et al. 2005 & 2011). Measurements of the abundance and
4
distribution of hydrogen in the coma can provide reliable estimates of water production rates in
comets (Mäkinen & Combi 2005, Combi et al. 2005 & 2011, Feldman et al. 2004). Observations
of 103P/Hartley 2 were made with the standard pipeline mode of daily full-sky observations.
SWAN has two sensor units, SU+Z and SU-Z, that typically observe north and south of
the ecliptic, respectively. The current sensitivity of SU+Z is such that an intensity of 1 Rayleigh
results in 0.24 counts per second per pixel. SU-Z is less sensitive than SU+Z by a factor of 2.6.
Each sensor has an instantaneous field of view (IFOV) of 5ox5o in a multi-anode detector of 25
1ox1o pixels. Images are made by mosaicking the IFOV across the sky in 2.5o increments.
Because of a spacecraft roll maneuver on October 29, 2010, all the observations of 103P/Hartley
2 were made with the SU+Z sensor even though the comet moved south of the ecliptic.
3. 103P/HARTLEY 2 OBSERVATIONS AND ANALYSIS
The full-sky SWAN images were examined beginning on August 1, 2010 for comet
103P/Hartley 2. The first image with a firm detection of the comet was the image on September
14, 2010. Thereafter, the comet was detected daily until October 16, when the comet was
obstructed by the SOHO spacecraft itself. The comet reemerged unobstructed on October 29 and
was observed on most days until December 12, 2010. On a number of days the comet was too
close to field stars, especially during the period from November 24 to 30, to enable a clean signal
from the comet to be isolated. Images of the position the comet on November 4 and November
10, 2010 are shown in Figure 1a-1b.
We used the model analysis procedure described for the SWAN observations of comet
1996 B2/Hyakutake (Mäkinen & Combi 2005). It combines the methods behind the syndyname
(Keller & Meier 1976) and the vectorial models (Festou 1981), while considering coma-wide
5
variations of input parameters and incorporating the necessary physical phenomena through the
inclusion of a parameterized (and less computationally intensive) version of the H atom velocity
distribution from Monte Carlo simulations (Combi & Smyth 1988) that account for the
expansion of the coma and partial thermalization of escaping H atoms.
Water production rates were calculated for each usable SWAN image from September 14
to December 12, 2010. The dissociation chain of water to OH radicals and the H atoms
produced, plus their transit times to fill the observable coma, introduces a time delay from any
change in water activity near the nucleus to an observable coma response of 1 to 2 days. The
large SWAN IFOV furthermore smears the significant short-term periodic variation produced by
the rotation of the nucleus (A'Hearn et al. 2011). The close geocentric distance of the comet
(0.11 to 0.35 AU) helps minimize this effect compared with previous some SWAN observations
of comets (Mäkinen & Combi 2005). Table 1 gives the observational circumstances as well as
the resulting water production for the SWAN observations. The g-factor is calculated from the
composite
site
LASP web
the
from
taken
data
Lyman-alpha
solar
http://lasp.colorado.edu/lisird/lya and the solar Lyman-α line profile by P. Lemaire et al. (1998).
Figure 1c-d shows the H coma as observed by SWAN a few hours after the EPOXI flyby, on
November 4, as well as a brightness profile cut through the coma.
Figure 2 shows the variation of the water production rate as a function time with the date
of the EPOXI flyby noted. The water production rate increased slowly from 1.5 to 2.5 x 1027
molecules s-1 from September 14 to 29 but then increased to 6.0 x 1027 molecules s-1 in 1 day.
Approaching perihelion it continued increasing until October 16 to a value of 8.7 x 1027
molecules s-1 at which time the location of the comet in the sky became obstructed by the SOHO
spacecraft. Then 3.7 days after perihelion the comet reemerged. The water production rate then
6
varied rather irregularly between 5.5 x 1027 and 1.2 x 1028 molecules s-1 until November 24 when
it was difficult to locate the comet among field stars again until November 30. By this time the
water production rate had dropped to a level slightly higher than the pre-September-30 level but
lower than the mean perihelion ±20-day level. The mean water production rate determined from
the SWAN observation on the day of the EPOXI flyby was 8.5 x 1027 molecules s-1.
4. RESULTS AND DISCUSSION
The water production rate variation from September 14 through September 24 shows no
evidence of correlation with the reported outburst of CN activity from the EPOXI team (A'Hearn
et al. 2011), but seems more similar to the variation of their dust-scattered continuum
observations, which increase rather monotonically throughout this period.
SWAN observed 103P/Hartley 2 during its 1997 apparition (Combi et al. 2011) yielding
water production rates that were consistent with values determined from both ground-based
observations of OH (A'Hearn et al. 1995) in 1991, O(1D) atoms (Fink 2009) in 1997 and from
observations with the ISOPHOT instrument on the Infrared Space Observatory (Crovisier et al.
1999; Colangeli et al. 1999). SWAN results from 1997 and 2010 are compared in Figure 2.
Clearly the comet and its production of water changed dramatically from 1997 (and 1991) to
2010. The 1997 production rates were a factor of 3 larger than those in the 2010 apparition.
One way to characterize water production rates in comets is to calculate an equivalent
surface area of water ice, which when exposed to sunlight at the comet's heliocentric distance, is
required to produce the observed water vapor. Because of the reality of variable surface and
surface fractional coverage by water this is called the "minimum active area," It was calculated
for all SWAN water production rates of 103P/Hartley 2 from 1997 and 2010 and compared with
7
the measured minimum, maximum and mean cross sections of the nucleus from EPOXI imaging
(A'Hearn et al. 2011), and all are plotted in Figure 3. The minimum active area is similar but
not equal to the active area. It is defined as A = LQr2/[NAFS(1-AV)], where L=50 kJ mol-1 is the
latent heat of water for sublimation, r is the heliocentric distance in AU, NA=6.022 x 1023 mol-1
(the Avogadro constant), FS=1365 W m-2 (the solar constant), and AV=0.03 (the assumed bond
albedo of the nucleus). See (Keller 1990) for a discussion of this definition.
For the majority of JFCs, the minimum active area calculated from the water production
rate is typically between 5 and 20% of the physical surface area of their nuclei. Such was the
case for the previous spacecraft flyby target comets: 1P/Halley, 19P/Borrelly, 81P/Wild 2 and
9P/Tempel 2 (A'Hearn et al. 1995; Fink 2009; Keller et al. 1987; Soderblom et al. 2002;
Brownlee et al. 2006). 1P/Halley, which is not a JFC, had the largest active fraction of these at
about 20%. The active area results for Hartley 2 are in many ways similar to those for long-
period Oort cloud comet 1996 B2/Hyakutake in that the minimum active area has been
comparable to or even larger than the projected cross section of the nucleus itself (πRN
2 ~ 1
million meters2). This was more so the case during the 1997 apparition of 103P/Hartley 2 when
the minimum active area was more than 3 times the mean projected cross section of the nucleus.
During outbursts of comet Hyakutake (Combi et al. 2005), the total production rate increased by
a factor of 4 above the 'normal' level, when many fragments were released from the nucleus,
including some large ones that were seen traveling down the tail for many days and producing an
extended source of gas (Harris et al. 1997; Desvoivres et al. 2000).
The EPOXI results (A'Hearn et al. 2011) show that the activity of the surface of
103P/Hartley 2 is not distributed uniformly over most of its surface. So even during the 2010
apparition, the fact that the minimum active area peaks at the value of the mean projected area of
8
the nucleus indicates that a significant fraction of its water production results from the extended
halo of icy fragments that appear to be carried off the surface by CO2-driven activity. The fact
that the activity was three times larger in 1997 means either that this process was far more
prevalent in 1997 (and also in 1991) or that some drastic alteration has occurred to the nucleus
since 1997, or both.
The synthesis of all the observations (EPOXI, space-based, and ground-based) of this
comet over the coming months and years will hopefully shed some light on a number of
fundamentally important issues of cometary science. Model analysis of observations of water
and its byproducts having a higher spatial resolution than the SWAN observations might be able
to quantitatively separate water production directly from the nucleus and that from the extended
cloud of fragments seen in the EPOXI images (A'Hearn et al. 2011) and shown to dominate the
global rate measured by SWAN. The role of highly volatile species (e.g., CO2) may have been
underestimated or at least underappreciated in the activity of comets, perhaps even inside the
solar-system's so-called snow line, challenging the current paradigms for cometary activity.
ACKNOWLEDGEMENTS
SOHO is an international cooperative mission between ESA and NASA. M. Combi
acknowledges support from grant NNG08AO44G from the NASA Planetary Astronomy
Program. J. -L. Bertaux, E. Quémerais and S. Ferron acknowledge support from CNRS and
CNES. J.T.T. Mäkinen was supported by the Finnish Meteorological Institute. We thank Dr.
M.F. A'Hearn for helpful discussions of preliminary EPOXI results prior to publication.
REFERENCES
9
A'Hearn, M. F., Millis, R. L., Schleicher, D. G., Osip, D. J., & Birch, P. V. 1995, Icarus, 118,
223
A'Hearn, M.F., & Combi, M.R. 2007, Icarus, 187, 1
A'Hearn, M.F. et al. 2011, Science, submitted
Bertaux, J.-L. et al. 1995, Solar Phys., 162, 403
Bertaux, J.-L., Kyrölä, E., Quemerais, E., Lallement, R., Schmidt, W., Summanen, T., Costa, J.,
& Mäkinen, T. 1999, Space Sci. Rev., 87, 129
Brownlee, D. et al. 2006, Science, 314, 1711
Colangeli, L. et al. 1999, A&A, 343, L87
Combi, M.R., Lee, Y., Pattel, T.S., Mäkinen, J.T.T., Bertaux, J.-L., & Quemérais, E. 2011, AJ, in
press
Combi, M.R., Mäkinen, J.T.T., Bertaux, J.-L., & Quemérais, E. 2005, Icarus, 177, 228
Combi, M.R., & Smyth, W.H. 1988, ApJ, 327, 1044
Crovisier, J. 1994, A&A, 213, 459
Crovisier, J. et al. 1999, ESA SP- 427, 161
Desvoivres, E. et al. 2000, Icarus, 144, 172
Fernández, J. 1980, MNRAS, 192, 481
Feldman, P.D., Cochran, A.L., & Combi, M.R. 2004, in Comets II, M.C. Festou, H.U. Keller,
H.A. Weaver, Eds., Univ. Arizona Press, pp. 425-447
Festou, M.C. 1981, A&A, 95, 69
Fink, U. 2009, Icarus, 301, 311
Hartley, M., 1986, IAU Circ. 4197, 1, Green, D.W.E., Ed.
Harris, W.M. et al. 1997, Science, 277, 676
10
Keller, H.U. 1990, in Physics and Chemistry of Comets, W.F. Huebner, Ed, Springer-Verlag,
Berlin, pp. 18-21
Keller, H.U. et al. 1987, A&A,187, 807
Keller, H.U., Meier, R.R. 1976, A&A, 52, 272
Lemaire, P., Emerich, C., Curdt, W., Schuehle, U., & Wilhelm, K. 1998, A&A, 334, 1095
Levison, H., & Duncan, M. 1997, Icarus, 127, 13
Lisse, C.M. et al. 2009, PASP, 121, 968
Mäkinen, J.T.T., Bertaux, J.-L., Combi, M.R., & Quémerais, E. 2001, Science, 292, 1326
Mäkinen, J.T.T., & Combi, M.R. 2005, Icarus, 177, 217
Soderblom, L. et al. 2002, Science, 296, 1087
11
Table 1. SOHO/SWAN Observations of Comet 103P/Hartley 2
and Water Production Rates in 2010
ΔT r Δ g Q ± δQ
(days) (AU) (AU) (s-1) (1027 s-1)
-42.853 1.211 0.285 0.002205 1.81 ± 1.0
-41.853 1.204 0.278 0.002192 2.29 ± 0.73
-40.853 1.198 0.271 0.002192 2.42 ± 0.65
-39.766 1.191 0.264 0.002192 2.07 ± 0.73
-38.766 1.185 0.257 0.002191 1.91 ± 0.74
-36.766 1.173 0.243 0.002191 3.56 ± 0.53
-34.766 1.162 0.230 0.002179 2.47 ± 0.55
-33.766 1.156 0.224 0.002178 2.37 ± 0.60
-32.765 1.151 0.217 0.002178 2.76 ± 0.73
-31.766 1.146 0.211 0.002167 2.39 ± 0.66
-30.765 1.141 0.205 0.002167 3.04 ± 0.75
-29.766 1.136 0.199 0.002167 2.05 ± 0.96
-28.765 1.131 0.193 0.002167 2.91 ± 0.77
-27.765 1.126 0.187 0.002167 2.52 ± 0.81
-26.765 1.122 0.181 0.002156 6.17 ± 0.42
-25.765 1.117 0.176 0.002156 6.07 ± 0.47
-24.740 1.113 0.170 0.002155 6.32 ± 0.32
-23.740 1.109 0.165 0.002155 6.66 ± 0.54
-22.740 1.105 0.160 0.002146 6.18 ± 0.53
-21.736 1.101 0.155 0.002145 7.65 ± 0.44
-20.736 1.097 0.150 0.002145 7.74 ± 0.34
-19.712 1.093 0.145 0.002136 6.65 ± 0.52
-18.712 1.090 0.141 0.002136 7.10 ± 0.82
-17.707 1.087 0.137 0.002136 7.56 ± 0.35
-16.706 1.084 0.133 0.002135 10.49 ± 0.49
-15.684 1.081 0.129 0.002127 8.00 ± 0.78
-14.677 1.078 0.126 0.002126 8.20 ± 0.27
-13.655 1.076 0.123 0.002126 8.04 ± 0.32
-12.655 1.073 0.120 0.002119 7.23 ± 0.42
-11.648 1.071 0.118 0.002119 8.91 ± 1.46
-10.627 1.069 0.116 0.002119 8.70 ± 0.38
3.704 1.060 0.134 0.002097 6.38 ± 0.12
4.713 1.061 0.138 0.002097 7.01 ± 0.45
5.733 1.062 0.142 0.002097 5.93 ± 0.14
7.742 1.064 0.150 0.002095 8.51 ± 0.63
8.762 1.066 0.154 0.002094 5.69 ± 0.33
9.762 1.067 0.159 0.002093 6.74 ± 0.28
10.771 1.069 0.164 0.002092 7.56 ± 0.08
13.791 1.076 0.178 0.002091 7.19 ± 0.12
14.791 1.078 0.183 0.002090 8.02 ± 0.10
12
15.800 1.081 0.188 0.002090 6.12 ± 0.13
18.650 1.090 0.203 0.002091 8.32 ± 0.10
19.651 1.093 0.208 0.002090 8.86 ± 0.12
20.651 1.097 0.213 0.002091 8.75 ± 0.10
21.660 1.101 0.219 0.002091 10.68 ± 0.08
22.659 1.104 0.224 0.002090 11.74 ± 0.08
23.660 1.108 0.229 0.002092 10.24 ± 0.10
24.659 1.112 0.235 0.002092 9.62 ± 0.11
25.659 1.117 0.240 0.002092 8.71 ± 0.10
26.680 1.121 0.246 0.002091 8.82 ± 0.11
33.688 1.156 0.284 0.002098 3.52 ± 0.36
36.608 1.172 0.300 0.002097 3.33 ± 0.17
37.608 1.178 0.305 0.002105 3.70 ± 0.29
38.608 1.184 0.311 0.002104 2.14 ± 0.48
39.608 1.190 0.316 0.002104 2.76 ± 0.52
40.608 1.196 0.322 0.002104 3.53 ± 0.34
41.608 1.203 0.327 0.002103 4.07 ± 0.27
42.608 1.209 0.333 0.002112 2.85 ± 0.35
43.608 1.215 0.339 0.002112 3.02 ± 0.38
44.607 1.222 0.344 0.002112 4.77 ± 0.27
45.608 1.229 0.350 0.002111 6.15 ± 0.23
ΔT: Time from perihelion on 2010 October 28.2570 UT in days
r : Heliocentric distance (AU)
Δ: Geocentric distance (AU)
g: Solar Lyman-α g-factor (photons s-1) at 1 AU
Q: Water production rates for each image (s-1)
δQ: internal 1-sigma uncertainties
13
Fig. 1. The sky in H Lyman-α as observed by the SWAN camera on SOHO. In (a) is the sky on
4 November 2010 the day of the EPOXI flyby and in (b) six days later. The comet is highlighted
with a black circle. These images are difference images with an image from November 1
subtracted from each. The noise that remains comes from incomplete subtraction of star images
owing to inexact spatial registration. The color scale on the right is in Rayleighs. In (c) is an
isolated image of the comet from November 4, 2010, projected with the sun to the right, and in
(d) is the intensity distribution along the red line cut shown in (c).
14
EPOXI Flyby
Fig. 2. Water production rate of comet 103P/Hartley 2. The triangles give the water production
rates determined from the SWAN H Lyman-α images from September 14 through December 12,
2010. The diamonds give the SWAN results from the 1997 apparition (Combi et al. 2011). The
vertical lines give the error bars due to internal sources of error only, namely instrument noise
and uncertainty due to the IPM subtraction. The day of the EPOXI flyby in 2010 is indicated.
There was a step function increase in water production of about a factor of 2.5 at T=-28 days
(October 1). There were data gaps between T=-12 and T=+2 (October 16 to 30) when SOHO
spacecraft obstructed the portion of the sky where the comet was located and then again from
T=+26 to T=+ 32 (24-30 November) when the comet was too close to nearby stars to get a clean
enough image to use. Some time during this period the comet seems to have gone through an
activity drop similar to the September 30 rise.
15
Maximum Cross Section
Mean Cross Section
Minimum Cross Section
Fig. 3. The Minimum Active Area (in m2) of comet 103P/Hartley 2 plotted as function of time
from perihelion. The triangles give the values calculated from the SWAN water production rates
in 2010 and the diamonds from 1997. The solid horizontal lines give the maximum and
minimum cross sections of the nucleus from EPOXI (A'Hearn et al. 2011) and the mean value
from Spitzer Space Telescope (Lisse et al. 2009). Since it is apparent from EPOXI resultsthat
the entire nucleus is not active, much of the water production seen within ±20 days of perihelion
must be due to the icy fragments released by the CO2-driven activity. Furthermore, since the
activity was 3 times higher in 1997, either the CO2-driven activity was much larger, or perhaps
some more drastic change happened to the nucleus since 1997.
16
|
1709.09959 | 1 | 1709 | 2017-09-26T18:00:01 | An Isolated Microlens Observed from K2, Spitzer and Earth | [
"astro-ph.EP",
"astro-ph.IM",
"astro-ph.SR"
] | We present the result of microlensing event MOA-2016-BLG-290, which received observations from the two-wheel Kepler (K2), Spitzer, as well as ground-based observatories. A joint analysis of data from K2 and the ground leads to two degenerate solutions of the lens mass and distance. This degeneracy is effectively broken once the (partial) Spitzer light curve is included. Altogether, the lens is found to be an extremely low-mass star located in the Galactic bulge. MOA-2016-BLG-290 is the first microlensing event for which we have signals from three well-separated ($\sim1$ AU) locations. It demonstrates the power of two-satellite microlensing experiment in reducing the ambiguity of lens properties, as pointed out independently by S. Refsdal and A. Gould several decades ago. | astro-ph.EP | astro-ph | Draft version July 5, 2021
Typeset using LATEX twocolumn style in AASTeX61
7
1
0
2
p
e
S
6
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
5
9
9
0
.
9
0
7
1
:
v
i
X
r
a
AN ISOLATED MICROLENS OBSERVED FROM K2, SPITZER AND EARTH
Wei Zhu (),1, 2 A. Udalski,3 C. X. Huang,4 S. Calchi Novati,5, 6 and T. Sumi7
--
R. Poleski,1, 3 J. Skowron,3 P. Mr´oz,3 M. K. Szyma´nski,3 I. Soszy´nski,3 P. Pietrukowicz,3 S. Koz(cid:32)lowski,3
K. Ulaczyk,3, 8 and M. Pawlak3
(OGLE Collaboration)
C. Beichman,9 G. Bryden,10 S. Carey,11 B. S. Gaudi,1 A. Gould,1, 12, 13 C. B. Henderson,10 Y. Shvartzvald,10, 14 and
J. C. Yee15
(Spitzer Team)
I. A. Bond,16 D. P. Bennett,17 D. Suzuki,18 N. J. Rattenbury,19 N. Koshimoto,7 F. Abe,20 Y. Asakura,20
R. K. Barry,17 A. Bhattacharya,17, 21 M. Donachie,19 P. Evans,19 A. Fukui,22 Y. Hirao,7 Y. Itow,20 K. Kawasaki,7
M. C. A. Li,19 C. H. Ling,16 K. Masuda,20 Y. Matsubara,20 S. Miyazaki,7 H. Munakata,20 Y. Muraki,20
M. Nagakane,7 K. Ohnishi,23 C. Ranc,17 To. Saito,24 A. Sharan,19 D. J. Sullivan,25 P. J. Tristram,26
T. Yamada,27 and A. Yonehara27
(MOA Collaboration)
1Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA
2Canadian Institute for Theoretical Astrophysics, 60 St George Street, University of Toronto, Toronto, ON M5S 3H8, Canada
3Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
4Department of Physics and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, Cambridge, MA
02139, USA
5IPAC, Mail Code 100-22, Caltech, 1200 E. California Blvd., Pasadena, CA 91125
6Dipartimento di Fisica "E. R. Caianiello", Universit`a di Salerno, Via Giovanni Paolo II, 84084 Fisciano (SA), Italy
7Department of Earth and Space Science, Graduate School of Science, Osaka University, 1-1 Machikaneyama, Toyonake, Osaka 560-0043,
Japan
8Department of Physics, University of Warwick, Gibbert Hill Road, Coventry, CV4 7AL, UK
9NASA Exoplanet Science Institute, MS 100-22, California Institute of Technology, Pasadena, CA 91125, USA
10Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA
11Spitzer Science Center, MS 220-6, California Institute of Technology, Pasadena, CA, US
12Korea Astronomy and Space Science Institute, 776 Daedeokdae-ro, Yuseong-Gu, Daejeon 34055, Korea
13Max-Planck-Institute for Astronomy, Konigstuhl 17, 69117 Heidelberg, Germany
14NASA Postdoctoral Program Fellow
15Smithsonian Astrophysical Observatory, 60 Garden St., Cambridge, MA 02138, USA
16Institute for Natural and Mathematical Sciences, Massey University, Private Bag 102904 North Shore Mail Centre, Auckland 0745, New
Zealand
17Laboratory for Exoplanets and Stellar Astrophysics, NASA/Goddard Space Flight Center, Greenbelt, MD 20771, USA
18Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1 Yoshinodai, Chuo, Sagamihara, Kanagawa
252-5210, Japan
19Department of Physics, University of Auckland, Private Bag 92019, Auckland, New Zealand
20Institute of Space-Earth Environmental Research, Nagoya University, Furo-cho, Chikusa, Nagoya, Aichi 464-8601, Japan
21Department of Physics, University of Notre Dame, Notre Dame, IN 46556, USA
22Okayama Astrophysical National Astronomical Observatory, 3037-5 Honjo, Kamogata, Asakuchi, Okayama 719-0232, Japan
Corresponding author: Wei Zhu
[email protected]
2
Zhu et al.
23Nagano National College of Technology, Nagano 381-8550, Japan
24Tokyo Metropolitan College of Industrial Technology, Tokyo 116-8523, Japan
25School of Chemical and Physical Sciences, Victoria University, Wellington, New Zealand
26University of Canterbury Mt John Observatory, PO Box 56, Lake Tekapo 7945, New Zealand
27Department of Physics, Faculty of Science, Kyoto Sangyo University, 603-8555 Kyoto, Japan
ABSTRACT
We present the result of microlensing event MOA-2016-BLG-290, which received observations from the two-wheel
Kepler (K2 ), Spitzer, as well as ground-based observatories. A joint analysis of data from K2 and the ground
leads to two degenerate solutions of the lens mass and distance. This degeneracy is effectively broken once the
(partial) Spitzer light curve is included. Altogether, the lens is found to be an extremely low-mass star located in the
Galactic bulge. MOA-2016-BLG-290 is the first microlensing event for which we have signals from three well-separated
(∼ 1 AU) locations. It demonstrates the power of two-satellite microlensing experiment in reducing the ambiguity of
lens properties, as pointed out independently by S. Refsdal and A. Gould several decades ago.
Keywords: gravitational lensing: micro -- methods: data analysis -- techniques: photometric --
parallaxes -- stars: fundamental parameters
Two-satellite Event MOA-2016-BLG-290
3
1. INTRODUCTION
The implementation of space-based microlensing par-
allax has revolutionalized the field of Galactic microlens-
ing (e.g., Dong et al. 2007; Udalski et al. 2015a). The
same microlensing event is seen to evolve differently in
views of ground-based telescopes and a space-based tele-
scope such as Spitzer or Kepler because of the large sep-
aration (∼ 1 AU) (Refsdal 1966; Gould 1994; Gould &
Horne 2013). This effect yields the microlensing paral-
lax vector πE, which conveys crucial information on the
lens mass and distance.
Although the Spitzer microlensing program has been
successful in terms of measuring masses of individual
planetary systems (Udalski et al. 2015a; Street et al.
2016; Shvartzvald et al. 2017; Ryu et al. 2017) and con-
straining the Galactic distribution of planets (Calchi No-
vati et al. 2015a; Yee et al. 2015b; Zhu et al. 2017a),
there is a generic uncertainty in measuring πE with a
single satellite, especially in cases of single-lens events.
The microlensing parallax vector πE is directly related
to the displacement between the two lens-source relative
trajectories
(cid:18) t0,sat − t0,⊕
tE
(cid:19)
πE ≈ AU
D⊥
, u0,sat − u0,⊕
,
(1)
where D⊥ is the separation between the satellite and
Earth perpendicular to the line of sight, t0 is the time
of maximum magnification, u0 is the impact parameter,
and subscripts "sat" and "⊕" denote those seen by the
satellite and Earth, respectively. Ambiguities arise be-
cause in majority cases, only u0 (rather than u0) can
be constrained by the light curve, thus leading to a four-
fold degeneracy in vector πE and a two-fold degeneracy
in its amplitude πE. Several studies have proposed ways
to break these degeneracies, and others pointed out spe-
cial situations in which such degeneracies do not matter
(see Yee et al. 2015a and references therein).
Along with proposing the idea of space-based mi-
crolensing parallax, Refsdal (1966) and Gould (1994)
also pointed out that the most efficient way to break
such parallax degeneracies should be to observe the same
microlensing event simultaneously from another well-
separated and misaligned location (satellite). The ad-
dition of a second satellite can effectively break the par-
allax degeneracies, especially the amplitude degeneracy.
Several decades after this idea was proposed, we finally
have the chance to test it. In 2016, the two-wheel Kepler
mission (K2, Howell et al. 2014) conducted a microlens-
ing campaign toward the Galactic bulge from April 22
to July 2, which overlapped with the Spitzer microlens-
ing campaign (June 18 to July 26) for nearly two weeks.
With this unique opportunity, a specific program (Gould
et al. 2015) was developed in order to demonstrate the
idea of Refsdal (1966) and Gould (1994). In total about
30 microlensing events received observations from both
satellites in addition to the dense coverage by ground-
based telescopes. 1 This work presents the first analysis
of this sample, specifically the bright single-lens event
MOA-2016-BLG-290 for which the microlensing signal
is detected from all three locations. 2
2. OBSERVATIONS & DATA REDUCTIONS
Microlensing event MOA-2016-BLG-290 was first
identified by the MOA (Microlensing Observations
in Astrophysics, Bond et al. 2001) collaboration at
UT 20:26 of 2016 June 1st (HJD(cid:48) =HJD−2450000 =
7541.35), based on observations from its 1.8-m telescope
with a 2.2 deg2 field at Mt. John, New Zealand. About
five days later, this event was also alerted as OGLE-
2016-BLG-0975 by the OGLE (Optical Gravitational
Lensing Experiment, Udalski et al. 2015b) Collabora-
tion through the Early Warning System (Udalski et al.
1994; Udalski 2003), based on data taken by the 1.3-m
Warsaw Telescope at the Las Campanas Observatory in
Chile.
With equatorial coordinates (R.A., decl.)2000 =
(18h04m57.s01,−28◦37(cid:48)40.(cid:48)(cid:48)1) and Galactic coordinates
(l, b)2000 = (2.◦40,−3.◦50), this event lies inside the mi-
crolensing super stamp of the K2 Campaign 9 (Gould
& Horne 2013; Henderson et al. 2016). It was therefore
monitored at 30 min cadence by K2 from 2016 April 22
to May 18 (C9a, HJD(cid:48) = 7501 − 7527) and from May
22 to July 2 (C9b, HJD(cid:48) = 7531 − 7572).
Events such as MOA-2016-BLG-290 that were ob-
served by K2 were preferentially selected during the
2016 Spitzer microlensing campaign, for the purpose
of demonstrating the two-satellite microlensing paral-
lax concept (Gould et al. 2015). In the current case, the
Spitzer team selected it as a Spitzer target subjectively
at UT 15:03 on 2016 June 9 (HJD(cid:48) = 7549.13), follow-
ing a revised protocol of Yee et al. (2015b). This selec-
tion turned into objective on June 12, meaning that this
event met our objective selection criteria. Because of
1 Although the two-satellite microlensing experiment with K2
and Spitzer is the first time that we observe the same event from
three well-separated locations, it is not the first time that one
event was observed by ground-based and two space-based tele-
scopes. In 2015, a few Spitzer microlensing targets were also ob-
served by Swift. See Shvartzvald et al. (2016) for the case of
OGLE-2015-BLG-1319.
2 The planetary event OGLE-2016-BLG-1190 presented in Ryu
et al. (2017) was also observed by K2, Spitzer, and ground-based
observatories, but the K2 data did not detect the microlensing
signal. Nevertheless, this non detection also led to the resolution
of the parallax degeneracy. See Ryu et al. (2017) for details.
4
Zhu et al.
Table 1. Best-fit parameters of microlensing models. With only
ground-based and K2 data, all four solutions are allowed, but only
the first two [(+, +) and (−,−)] survive once Spitzer data and the
the associated source I − [3.6µm] color constraint are taken into
account.
Parameters
t0,⊕ − 7552
u0,⊕
tE
πE,N
πE,E
I − [3.6µm]
(+, +)
0.380(4)
0.7032(6)
6.370(8)
0.180(5)
−0.150(4)
−5.29(13)
(−, −)
(+, −)
(−, +)
0.378(4)
−0.7033(6)
6.370(9)
−0.200(5)
−0.131(4)
−5.35(13)
0.375(4)
0.7036(4)
6.379(9)
−2.199(6)
−0.139(4)
−4.89(13)
0.383(4)
−0.7031(6)
6.373(9)
2.156(4)
−0.370(4)
−4.53(13)
the Sun-angle limit, the Spitzer observations were taken
between HJD(cid:48) = 7559.6 and 7571.2 at a quasi-daily ca-
dence. All Spitzer observations were taken in the 3.6 µm
channel.
The ground-based data were reduced using the stan-
dard or variant version of the image subtraction method
(Alard & Lupton 1998; Wozniak 2000; Bramich 2008).
The raw K2 light curve was extracted and modeled fol-
lowing the method of Zhu et al. (2017b), which is a
special application of Soares-Furtado et al. (2017) and
Huang et al. (2015). The Spitzer data were reduced us-
ing the software that was customized for the microlens-
ing program (Calchi Novati et al. 2015b).
3. BREAKING PARALLAX DEGENERACY WITH
TWO-SATELLITE EXPERIMENT
First, as a proof of concept of the two-satellite mi-
crolensing parallax method, we choose to only model the
ground-based and K2 data, and then compare the pre-
dicted Spitzer light curve with the actual Spitzer data.
The modeling of ground-based and K2 data follows
the methodology of Zhu et al. (2017b), but with a minor
modification. According to the OGLE-III Catalog of
Variable Stars (Soszy´nski et al. 2013), a low-amplitude
(0.034 mag) long-period (342.5 days) variable, OGLE-
BLG-LPV-202211, sits only 10(cid:48)(cid:48) (or 2.5 K2 pixels) away
from the location of MOA-2016-BLG-290. Due to the
broad point spread function and the unstable pointing of
the K2 spacecraft, this variable star affects the raw K2
light curve that we extracted. To minimize the influence
of this nearby variable star, we introduce an additional
term that scales linearly with time into the model of
K2 raw light curve (Equation 1 of Zhu et al. 2017b).
Following Zhu et al. (2017b), we also include a constraint
on the source Kp − I color, which is derived from the
source V − I color from OGLE photometry.
With only data from the ground and from K2 included
in the modeling, the four-fold degeneracy emerges.
These are generally denoted as (Earth-K2) (±,±) solu-
tions, with the first sign and the second sign indicating
the sign of u0,⊕ and u0,K2 in the geocentric frame,
respectively (Zhu et al. 2015). The microlensing param-
eters for these solutions are given in Table 1, and the
microlensing geometries are shown in Figures 1 and 2
for different choices of reference frame. As these plots
illustrate, the four solutions are distinct in terms of the
velocity vector of the lens-source relative motion, which
is directly determined by the parallax vector πE. If only
the amplitude of parallax is considered, the four-fold de-
generacy essentially collapses to two-fold (Gould 1994),
which we denote as "small πE" [(+, +) and (−,−)] and
"large πE" [(+,−) and (−, +)].
Because Spitzer 's position relative to Earth and K2
is well known, we can then "predict" the microlensing
light curve that Spitzer would see for all four solutions.
These predicted Spitzer light curves are shown in the
left panel of Figure 3 together with the ground-based
(OGLE & MOA) and K2 light curves. Given the dif-
ferent behaviors of the predicted light curves, Spitzer
observations would in principle pick out the correct so-
lution if this third observer had a full coverage of the
event light curve. Unfortunately, Spitzer was not able
to observe this event until HJD(cid:48) = 7559.5 because of
its Sun-angle limit, and therefore it only captured the
falling tail of the light curve. Such a partial light curve
can be fit by all four solutions equally well, if no other
information is provided.
Fortunately, with the known properties of the source
star, we are able to at least break the degeneracy in
the parallax amplitude πE (i.e., "small πE" vs. "large
πE"). The Spitzer (as well as OGLE and MOA) flux is
modeled by
F (t) = FS · A(t) + FB .
(2)
Here FS is the source flux in given observatory/bandpass,
FB is the flux that is within the aperture but unrelevant
to the microlensing effect, and A(t) is the microlensing
magnification at given time t. For the same set of Spitzer
measurements, a different magnification behavior would
suggest a different source brightness FS (and so source
color I − [3.6µm], since the source I magnitude is well
determined). With the predicted Spitzer light curves
from the previous step, the source I − [3.6µm] colors
are then estimated for all four solutions, and they are
listed in Table 1 as well. The uncertainty on the source
color is dominated by uncertainties on the Spitzer ob-
servations. For the two groups of solutions, the inferred
Two-satellite Event MOA-2016-BLG-290
5
Figure 1. Trajectories of three observers (Earth, K2, and Spitzer ) with respect to the aligned lens and source position (marked
as asterisks). These are the geocentric views of microlensing geometries (Gould 2004). The four solutions allowed by the ground-
based and K2 data are shown individually, and the predicted Spitzer positions are shown in red, with the solid line denoting
the time span of actual Spitzer observations. For each trajectory, there are three arrows indicating the direction of motion at
three different epochs: HJD(cid:48) = 7540, 7550, and 7560, respectively. The trajectories are oriented so that north is up and east is
left (see Skowron et al. 2011 for the sign convention of u0). We note that the Earth-K2 -Spitzer relative positions (in AU) are
the same in all plots, and that they are simply scaled differently in all plots for the given parallax measurements.
source I − [3.6µm] colors are statistically different at
> 3 σ level.
πE" solutions are allowed. See Figure 4 for the illustra-
tion of the color determination and comparison.
We then derive the source color
I − [3.6µm] = −5.56 ± 0.12
(3)
from a model-independent way, by substituting the
source V − I color into the stellar I − [3.6µm] vs. V − I
color-color relation. This relation is established based
on neighboring field stars with similar properties (see
details in Calchi Novati et al. 2015b). The deviations
between this color measurement and inferred colors are
1.5σ, 1.2σ, 3.8σ, and 5.8σ for (+, +), (−,−), (+,−), and
(−, +) solutions, respectively. Therefore, the "large πE"
solutions can be securely rejected, and only the "small
As a final step, we model data from all observatories
(OGLE, MOA, K2 and Spitzer ) simultaneously. The
microlensing parameters are almost identical to those
for Earth-K2 (+, +) and (−,−) solutions, and therefore
are not listed separately here. The best-fit models and
all data sets are illustrated in the right panel of Figure 3.
4. DISCUSSION
In this work, we present the analysis of the first mi-
crolensing event that has detected signals from at least
three well-separated (∼ 1 AU) locations, which in the
current case are Earth, K2, and Spitzer. With data
from the third well-separated observer, we have demon-
202North (rE)EarthK2SpitzerEarth-K2 (+,+)(,)42024East (rE)202North (rE)(+,)42024East (rE)(,+)6
Zhu et al.
Figure 2. Trajectories of the lens with respect to the aligned source and observer (Earth at t0,⊕), and the motions of all three
observers relative to the same reference point. These are the heliocentric views of microlensing geometries (Calchi Novati &
Scarpetta 2016). Again, north is up and east is left, but we use the same physical scale for all solutions so that the motions of
observers are the same. For each curve, the arrow indicates the position of the object as well as its directory of motion. Now
the Earth-trailing orbits of K2 and Spitzer are clearly seen.
Figure 3. Left panel: predicted Spitzer light curves (in red) based on the modeling of the ground-based (OGLE in black and
MOA in orange) and K2 C9 data (in blue). The raw K2 data are shown as blue open dots, and the binned data are shown
as blue solid dots with error bars. The four predicted Spitzer light curves are plotted with different line styles: solid, dashed,
dash-dot, and dotted for Earth-K2 (+, +), (−,−), (+,−), and (−, +) solutions, respectively. Right panel: Data and the best-fit
models for all observatories. Here we only show the re-binned K2 C9 data, and the remaining labels are the same as those in the
left panel. Once all data and the color constraints are included, only the "small πE" solutions, (+, +) and (−,−), are allowed.
1812606East (AU)1260612North (AU)(+,+)(,)1.60.80.00.81.6East (AU)1.60.80.00.81.6EarthK2SpitzerLens(+,)(,+)3040506070HJD-24575001.01.21.41.6MagnificationOGLEMOAK2C9Spitzer (predicted)3040506070HJD-24575001.01.21.41.6MagnificationOGLEMOAK2C9 (binned)SpitzerTwo-satellite Event MOA-2016-BLG-290
7
the near disk (DL = 2.5± 0.8 kpc). The ability to break
the πE amplitude degeneracy really reduces the uncer-
tainties on the lens properties.
Among other proposed methods (Gould 1995, 1999;
Gould & Yee 2012; Yee et al. 2015a; Calchi Novati et
al. 2015a), obtaining observations from a third location
has always been considered the most efficient in breaking
the parallax degeneracy in satellite microlensing exper-
iments. It is nevertheless difficult to do so for a large
number of events because of obvious economical reasons.
In the absence of this method, the Rich argument can
be applied for statistical purposes (Calchi Novati et al.
2015a; Zhu et al. 2017a). In fact, for the present case,
the application of the Rich argument would also argue
for an extremely low probability (1%) of the "large πE"
solutions. However, the resolution of degeneracy is im-
portant whenever precise knowledge is required of an
individual event. Therefore, the ∼30 events observed in
the 2016 two-satellite microlensing experiment are very
valuable. This ensemble can be used to refine the mi-
crolensing parallax method, which is the foundation of
the on-going Spitzer microlensing project and will likely
be a crucial component of the future space-based mi-
crolensing surveys.
Regardless, one may wonder when will be the next
time to apply this
two-satellite parallax method?
Within the predictable period, one could only hope
to apply this method one decade from now to the Wide
Field InfaRed Survey Telescope (WFIRST, Spergel et al.
2015) and Euclid (Penny et al. 2013), although the sit-
uation will be largely different because these telescopes
will likely be at Earth-Sun L2 point (i.e., a much shorter
baseline, see Zhu & Gould 2016 for a detailed analysis
of WFIRST parallax. See also Yee 2013). Nevertheless,
the history of microlensing parallax has already proved
that fantastic scientific ideas will never be buried.
Work by W.Z. and A.G. were supported by NSF grant
AST-1516842. Work by S.C.N. and A.G. were sup-
ported by JPL grant 1500811. R.P. acknowledges sup-
port from K2 Guest Observer program under NASA
grant NNX17AF72G. Work by Y.S. was supported by an
appointment to the NASA Postdoctoral Program at the
Jet Propulsion Laboratory, California Institute of Tech-
nology, administered by Universities Space Research As-
sociation through a contract with NASA. Work by C.R.
was supported by an appointment to the NASA Post-
doctoral Program at the Goddard Space Flight Center,
administered by USRA through a contract with NASA.
This paper includes data collected by the Kepler mis-
sion. Funding for the Kepler mission is provided by the
NASA Science Mission directorate. Some of the data
Figure 4. This figure shows the stellar I − [3.6µm] vs. V − I
color-color relation as well as the data points used to derive
it. The shaded region remarks the 1-σ uncertainty of this
color-color relation. The open squares are the source colors
inferred from the four solutions. The colors of the "small πE"
solutions (+ + & − −) are consistent with this independent
color measurement, while those of the "large πE" solutions
(+ − & − +) are inconsistent at > 3σ level.
strated that the generic parallax degeneracy arising in
the single-satellite microlensing parallax experiment is
effectively broken. This is essentially the first realiza-
tion of the decades-old idea proposed by Refsdal (1966)
and Gould (1994) independently.
For the current event, we could only break the degen-
eracy in the (two-fold) parallax amplitude rather than
the degeneracy in the (four-fold) parallax vector. This
is partly because this event could not be observed by
Spitzer until it was almost finished, but mostly because
the event is near the ecliptic plane and the Earth-K2 -
Spitzer configuration is nearly colinear (Gaudi & Gould
1997). Nevertheless, it is the amplitude of the microlens-
ing parallax that matters in determining the lens prop-
erties, and therefore being able to break the degeneracy
in parallax amplitude has already enabled a better de-
termination of the lens properties, such as lens mass. We
use the current case as an example. Using the Galactic
model and the Bayesian method of Zhu et al. (2017a),
we estimate the lens of MOA-2016-BLG-290 has a mass
ML = 77+34−23 MJ and distance DL = 6.8 ± 0.4 kpc for
accepted solutions. These values correspond to a brown
dwarf or extremely low-mass star likely in the Galactic
bulge. However, the other two solutions, had they been
correct, would suggest a high-mass (7+4−3 MJ) planet in
1.51.61.71.81.92.0VI6.56.05.55.04.5I[3.6m](+,+)(,)(+,)(,+)8
Zhu et al.
presented in this paper were obtained from the Mikul-
ski Archive for Space Telescopes (MAST). STScI is op-
erated by the Association of Universities for Research
in Astronomy, Inc., under NASA contract NAS5-26555.
Support for MAST for non-HST data is provided by the
NASA Office of Space Science via grant NNX09AF08G
and by other grants and contracts. OGLE project
has received funding from the National Science Cen-
tre, Poland, grant MAESTRO 2014/14/A/ST9/00121
to A.U.. The MOA project is supported by JSPS
Kakenhi grants JP24253004, JP26247023, JP16H06287,
JP23340064 and JP15H00781 and by the Royal Society
of New Zealand Marsden Grant MAU1104.
REFERENCES
Alard, C., & Lupton, R. H. 1998, ApJ, 503, 325
Bond, I. A., Abe, F., Dodd, R. J., et al. 2001, MNRAS,
327, 868
Bramich, D. M. 2008, MNRAS, 386, L77
Calchi Novati, S., Gould, A., Udalski, A., et al. 2015, ApJ,
804, 20
Calchi Novati, S., Gould, A., Yee, J. C., et al. 2015, ApJ,
814, 92
Calchi Novati, S., & Scarpetta, G. 2016, ApJ, 824, 109
Dong, S., Udalski, A., Gould, A., et al. 2007, ApJ, 664, 862
Gaudi, B. S., & Gould, A. 1997, ApJ, 477, 152
Gould, A. 1994, ApJL, 421, L75
Gould, A. 1995, ApJL, 441, L21
Gould, A. 1999, ApJ, 514, 869
Gould, A. 2004, ApJ, 606, 319
Gould, A., & Yee, J. C. 2012, ApJL, 755, L17
Gould, A., & Horne, K. 2013, ApJL, 779, L28
Gould, A., Yee, J., & Carey, S. 2015, Spitzer Proposal,
12015
Henderson, C. B., Poleski, R., Penny, M., et al. 2016,
PASP, 128, 124401
Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126,
398
Huang, C. X., Penev, K., Hartman, J. D., et al. 2015,
MNRAS, 454, 4159
Penny, M. T., Kerins, E., Rattenbury, N., et al. 2013,
MNRAS, 434, 2
Refsdal, S. 1966, MNRAS, 134, 315
Ryu, Y.-H., et al. 2017 ApJ submitted
Shvartzvald, Y., Li, Z., Udalski, A., et al. 2016, ApJ, 831,
183
Shvartzvald, Y., Yee, J. C., Calchi Novati, S., et al. 2017,
ApJL, 840, L3
Skowron, J., Udalski, A., Gould, A., et al. 2011, ApJ, 738,
87
Soares-Furtado, M., Hartman, J. D., Bakos, G. ´A., et al.
2017, PASP, 129, 044501
Soszy´nski, I., Udalski, A., Szyma´nski, M. K., et al. 2013,
AcA, 63, 21
Spergel, D., Gehrels, N., Baltay, C., et al. 2015,
arXiv:1503.03757
Street, R. A., Udalski, A., Calchi Novati, S., et al. 2016,
ApJ, 819, 93
Udalski, A., Szymanski, M., Kaluzny, J., et al. 1994, AcA,
44, 227
Udalski, A. 2003, AcA, 53, 291
Udalski, A., Yee, J. C., Gould, A., et al. 2015, ApJ, 799, 237
Udalski, A., Szyma´nski, M. K., & Szyma´nski, G. 2015,
AcA, 65, 1
Wozniak, P. R. 2000, AcA, 50, 421
Yee, J. C. 2013, ApJL, 770, L31
Yee, J. C., Udalski, A., Calchi Novati, S., et al. 2015, ApJ,
802, 76
Yee, J. C., Gould, A., Beichman, C., et al. 2015, ApJ, 810,
155
Zhu, W., & Gould, A. 2016, Journal of Korean
Astronomical Society, 49, 93
Zhu, W., Udalski, A., Gould, A., et al. 2015, ApJ, 805, 8
Zhu, W., Udalski, A., Calchi Novati, S., et al. 2017,
arXiv:1701.05191
Zhu, W., Huang, C. X., Udalski, A., et al. 2017, PASP, 129,
104501
|
1605.08471 | 2 | 1605 | 2016-11-06T23:29:36 | Do planetary seasons play a fundamental role in attaining habitable climates? | [
"astro-ph.EP",
"physics.ao-ph"
] | A simple phenomenological account for planetary climate instabilities is presented. The description is based on the standard model where the balance of incoming stellar radiation and outward thermal radiation is described by the effective planet temperature. Often, it is found to have three different points, or temperatures, where the influx of radiation is balanced with the out-flux, even with conserved boundary conditions. Two of these points are relatively long-term stable, namely the point corresponding to a frozen-climate and the point corresponding to a hot-climate. The hypothesis promoted in this paper is the possibility that it is the intermediate third point which is the basis for habitable-climates. I.e. that this initially unstable point is made relatively stable over a long period by the presence of seasonal climate variations. This points to the axial inclination, and perhaps the presence of orbital eccentricity, as the origin of the stability of the habitable point. An analysis involving the inclination, the size of the ice caps, and the length of the year shows that within the currently accepted value of the heat capacity of the Earth, the otherwise unstable habitable point is stabilized. | astro-ph.EP | astro-ph | arXiv version
6
1
0
2
v
o
N
6
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
7
4
8
0
.
5
0
6
1
:
v
i
X
r
a
Do planetary seasons play a fundamental role in
attaining habitable climates?
Kasper W. Olsen(a) and Jakob Bohr
DTU Nanotech, Building 344, Ørsteds Plads, Technical University of Denmark, DK-2800
Kongens Lyngby, Denmark
PACS 91.45.Nc -- Evolution of the Earth
PACS 92.70.Aa -- Abrupt/rapid change in climate
PACS 96.
PACS 97.82.-j -- Extrasolar planetary systems
-- Solar systems; planetology
Abstract -- A simple phenomenological account for planetary climate instabilities is presented.
The description is based on the standard model where the balance of incoming stellar radiation and
outward thermal radiation is described by the effective planet temperature. Often, it is found to
have three different points, or temperatures, where the influx of radiation is balanced with the out-
flux, even with conserved boundary conditions. Two of these points are relatively long-term stable,
namely the point corresponding to a frozen-climate and the point corresponding to a hot-climate.
The hypothesis promoted in this paper is the possibility that it is the intermediate third point
which is the basis for habitable-climates. I.e. that this initially unstable point is made relatively
stable over a long period by the presence of seasonal climate variations. This points to the axial
inclination, and perhaps the presence of orbital eccentricity, as the origin of the stability of the
habitable point. An analysis involving the inclination, the size of the ice caps, and the length of the
year shows that within the currently accepted value of the heat capacity of the Earth, the otherwise
unstable habitable point is stabilized.
Introduction. -- Since the advent of the first modern climate models in the 1960's
[1, 2], the past climate of the Earth has been a subject of intense debate. One possibility
is that at a point in the past the Earth has been fully, or at least nearly fully, covered by
ice [3, 4]. This is the so-called Snowball Earth hypothesis. The scientific debate continues
as to whether this hypothesis is true or not [5, 6]. Regardless of the final outcome of this
debate, it is worthwhile to notice that the debate would not have set off if it was not for the
pioneering insight that the balance between incoming heat and outgoing heat can have more
than one solution. This revelation has been an early stimulation for the work on current
climate models of the Earth [7]. See also data on radiative forcing and albedo feedback [8].
Recent work also considers the effects of daily cycles (diurnal asymmetry) [9].
In this paper we discuss the nature of the planetary heat balance in its most simple and
conceptual form. The results therefore apply to planet Earth, as well as to exoplanets in the
habitable zone [10]. We shall revisit the question of which of the possible solutions to the heat
balance equation correspond to a habitable climate. In the case of three such points we will
designate them the planetary snowball point, the habitable point, and the desert planet point
(a)Email: [email protected]
p-1
K.W. Olsen and J. Bohr
by increasing temperature. Hence, what we suggest is a reinterpretation, the suggestion is
that what is normally considered the unstable point is actually the habitable point. This
assignment follows from the "albedo curve" taking habitable to be in the coexistence regime
of water and ice. At first, it may seem that only the snowball point and the desert point
are stable equilibrium points [7]. This interpretation is based on a static analysis of how
the two power terms develop when one moves to a point slightly away from the equilibrium
points.
In general, the way one consider dynamical systems was pioneered by Lyapunov
[11] and has later led to a significant impact on the understanding of dynamical systems,
attractors, and Lyapunov exponents [12, 13]. One crucial observation is that under certain
conditions forced oscillations can stabilize an otherwise unstable equilibrium, e.g. as in the
dynamically driven hill-top stability [14].
We suggest that the habitable point can locally become stabilized for a long extended
period of time by the climate fluctuations which arise from seasonal changes that are caused
by the axial inclination of the planet. I.e. that seasons play an important role for climate
stability. A simplified stability analysis suggest that a significant axial inclination is needed.
Method. -- Following the standard model [7], a heuristic model that demonstrates the
kind of stability one can obtain for the heat balance can be developed. The planetary heat
radiation is described by an effective temperature TE such that the total re-radiated power,
PO, is given by the well-known Stefan-Boltzmann law,
PO = KST 4
E ,
(1)
where the constant KS is given by the product of the surface area of the planet, A, the emis-
sivity, , and Stefan-Boltzmann's constant, σ. Note that PO is monotonously increasing as
a function of the effective temperature with an everywhere positive second order derivative.
The expression for the total received power, PI , involves the planetary "whiteness" i.e.
albedo, and is assumed to depend on the effective temperature TE. We have,
PI = (1 − α)PS ,
(2)
where PS is the incoming power, e.g. stellar or solar radiation, and α the effective albedo.
The latter is defined as the fraction of the received power which is re-radiated without prior
absorption. We shall assume α ∈]0, 1[. Many simple functional forms of the albedo with
temperature, e.g. being a constant α(TE) = α0, or having an everywhere negative curvature
leaves only one solution for the requirement of heat balance, i.e.
PI = PO .
(3)
More complicated forms for the albedo as a function of temperature will have one or a
higher odd number of equilibrium solutions -- an even number of equilibrium points requires
touching but non-crossing lines. In the general case, we have to solve the equation
(1 − α(TE))PS = KST 4
E .
(4)
Below, we shall study the case of three equilibrium points which can arise when the albedo
function is loosely speaking an elongated Z-shape consisting of three straight pieces: At low
temperature a horizontal line corresponding to the single phase of ice, at high temperature
a horizontal line for the single phase of water, and finally an inclined line at intermediate
temperatures representing coexisting phases, see Figure 1A. Sometimes an S-shaped function
is used. For simplicity, we can construct both scenarios using a sigmoid-like function fi,
where
f0(x) = (1 + e−x)−1 ,
p-2
(5)
Seasons and habitable climates
and
0
x + 1/2
1
f1(x) =
if x < −1/2
if −1/2 ≤ x ≤ 1/2
otherwise
The "albedo curve" is then chosen to be of the form,
1 − α(TE) = α0 + (α∞ − α0)fi((TE − T0)/w) .
(6)
(7)
For i = 0 the standard S-shaped logistic function is recovered; for i = 1 one obtains the
elongated Z-shape. The constant w is proportional to the inverse of the slope of the albedo
curve at T = T0. Any other sigmoid -- like function, i.e. a monotonic function that interpo-
lates between 0 and 1, with slope equal to one at x = 0, could have been used with the same
qualitative, though not quantitative, results. The motivation for introducing this functional
form for the albedo curve is the well-known scenario used in the snowball Earth hypothesis,
namely, that the albedo of water can change drastically with its thermodynamic phase and
hence can change continuously when the effective temperature allows for the coexistence of
phases. Typically, this means an effective temperature which is not too far from the triple
point (T0 = 273.16K).
Let us assume that we have the situation with three crossings, i.e. three equilibrium
temperatures, as depicted in Figure 1A. The three points are denoted the S, the H, and the
D points from the associations snowball planet, habitable planet, and desert planet. The
corresponding effective temperatures are denoted {TS, TH , TD}. Both of the two dominantly
single phase points, S and D, are in a stable equilibrium. The reason is that the Taylor
expansion of P = PI − PO at these points have a negative first order derivative,
∂T P =
∂P (T ∗)
∂T
< 0 , where T ∗ ∈ {TS, TD} .
From this, we observe that with a small positive temperature change, ∆T ,
P (T ∗ + ∆T ) (cid:39) ∂T P ∆T < 0 ,
(8)
(9)
therefore the planet is cooled, and hence the system is driven towards the stable equilibrium
point, T ∗. This is not the case for the habitable point, H. At this multiphase point the first
order Taylor coefficient ∂T P is positive and any deviation from equilibrium will be amplified
over time. Hence, this is an unstable equilibrium solution. Can this unstable point exhibit
relatively long-term stability?
Let us consider the axial inclination of a planet which has icecaps at the two poles (in
our model, we assume that the icecaps are of the same size, see Figure 2B). This means that
the incoming heat varies with the seasons of the year, i.e. it depends on how the planetary
rotation axis is oriented relative the Sun or star. If we only consider the first order expansion
of P , the integral of the incoming heat will be linear and the result will again be instability.
It is necessary to consider the second and third order Taylor coefficients,
P (T + ∆T ) = P (T ) + ∂T P ∆T +
1
2
∂2
T P (∆T )2 +
1
6
∂3
T P (∆T )3 + . . .
(10)
The second order coefficient can help to form stability half of the year but causes a similar
instability the other half of the year. What about the third order coefficient? Akin to the
first- and second-order terms, neither the third- nor the higher-order terms are sufficient to
create stability. Therefore, a time-dynamic effect is needed.
One time-dynamic effect is the presence of seasonal albedo variations. Other factors
which can contribute to periodic changes in the heat balance such as eccentricity of orbits are
ignored and the planet is assumed perfectly spherical for simplicity. The received radiation
at an area element dA is:
p-3
K.W. Olsen and J. Bohr
Fig. 1: Energy balance of the Earth: (A) received power PI , here given per area, and
radiated power PO, also per area, as a function of effective temperature. PI is proportional
to 1 minus the albedo, which at low and at high temperature corresponds respectively to
complete ice cover, and no ice cover; (B) zoom of the energy balance, the three equilibrium
temperatures are ∼ 240K, 280K and 296K. An effective greenhouse factor of 0.55 was used
in the calculations (see [7]) and w = 10 was used for the sigmoid function. The albedo is
taken to be in the range, α ∈ [0.3, 0.7].
dPI (θ, φ, Ψ, Ω) = (1 − α(θ)) H( (cid:126)R(θ, φ, Ψ, Ω) · (cid:126)S) S0 dA ,
(11)
where (cid:126)R is the normal vector to the area element, (cid:126)S is the direction of the Sun or star, and S0
the solar constant. H is the Heaviside step function. The variable θ describes the planetary
latitude while φ describes the longitude. The albedo is assumed to be rotationally symmetric
and hence only depends on θ. The tilt of the planetary axis (the obliquity) is denoted Ψ
and the angle Ω tracks the phase of the planetary position in its orbit. The average value
of the received radiation can now be found by integrating over θ, φ and averaging over Ω (a
year), as Ψ is perceived to be constant. For the Earth the current obliquity is Ψ = 23.4◦.
The seasonal variation of the absorbed power is found by avoiding to average over Ω,
see Figure 2A. This figure shows the two extrema, equinox versus summer/winter solstice.
For ice caps of size 20◦, the difference in fractional area, and therefore total power, is seen
to be about 2%. Utilizing this observation, one can estimate the time constant related to
seasonal changes (it is not simply one year!), and compare it to the time constant related
to the instability of the habitable point: The effective global heat capacity of the Earth has
been estimated in [21, 22] to be C = 21.8 ± 2.1 W yr m−2 K−1. The derivate of the total
power with respect to temperature, λj = (dP/dT )j, determines the time constant at an
equilibrium point, as
τj = C/λj .
(12)
Using Figure 1, the three equilibrium points therefore have characteristic times constants
of τS = 12(±1), τD = 7(±0.6), τH = −84(±8) yr. The first two of these are positive,
while the one for the habitable point is negative (since it is an unstable point). However,
the unstable point H is also subject to seasonal variations. On the short time-scale, the
seasonal variations correspond to a time constant τΨ. This time constant τΨ is determined
by the condition, that 1− 2% = 0.98 = exp{−(1/4 yr)/τΨ}, i.e. τΨ (cid:39) 12 yr. For the seasons
to counter the habitable-point instability, we need approximately τH > 2τΨ. As 84 (±8)
years is greater than 2 × 12 years, this is fulfilled. However, if we imagine that the seasonal
effect was only three times smaller, then we would get 2× τΨ = 75 yr, and we could quickly
p-4
Temperature [K]050100150200250300350Power [Wm-2]050100150200250300350400450500PIPOTemperature [K]230240250260270280290300Power [Wm-2]80100120140160180200220240260PIPObe in trouble. In other words, the amount of axial inclination, the size of the ice caps and
the orbital period are all crucial in having a stable climate.
Seasons and habitable climates
Fig. 2: (A) Fraction of the projected surface area with minimal albedo value (α = 0.3) as
a function of the size of the two ice caps given in degrees -- at 90◦ the planet is completely
covered in ice. The upper curve (brown) is calculated at equinox, the lower (blue) curve at
summer/winter solstice. The small dent at 45◦ is a numerical inaccuracy. (B) Our idealized
Earth with an inclination of 23.4◦. The ice caps are here illustrated with a size of 30◦, the
present size is about 20◦.
Discussion. -- The suggested generic assignment of the habitable point deviate from
previous assignments, where it has been suggested that the point which we call D is the
habitable point [7]. The rationale behind our chosen assignment can best be seen if the
albedo function is taken in its most simple form, as in Figure 1A, corresponding to ice
phase, co-existing phases and hot phase respectively.
The habitable point H is an unstable point in the static situation. However, because of
the inclination of the axis of the Earth, it is a strongly driven system which leads to the
seasonal changes of the climate: The heat capacity of the Earth leads to a delay in establish-
ing equilibrium and hence a corresponding non-linearity. Indeed, strongly driven systems
with forced oscillations can have stable equilibria [14, 15]. Other papers have discussed the
possibility of attaining stability by perturbations [16, 17].
Previously, there has been a discussion about whether eccentricity and planetary pre-
cession would lead to a planet temporary leaving the habitable zone, or otherwise having
too extreme seasonal changes to support life as we know it from Earth [18 -- 20]. These are
valid concerns to be considered, yet at the same time one needs to incorporate the findings
of the present work, namely that we shall treasure our seasons as it is their very existence
that seem to stabilize the habitable climate.
Our suggestion can also be important in the search for habitable exoplanets as many
planets do not have a large angle of inclination. This will limit the number of observed
exoplanets that will be candidates for habitation. However, we believe, that when searching
long enough planets in the right belt with sufficiently inclined axis will be found to display
a similar intermediate stability.
Finally, the stability analysis of equilibrium points has implications for the on-going
climate debate. It becomes evident that one cannot in all cases obtain a new and modi-
fied climate by smoothly adjusting the boundary conditions, as for example the amount of
Greenhouse effect. The reason is that if the climate is modified beyond a certain point, a
far instability will arrive and the planet is forced into either the desert planet point or into
p-5
23.4K.W. Olsen and J. Bohr
the snowball point, neither of which will be a gentle climate for the human species.
REFERENCES
[1] Budyko M I, The effect of solar radiation on the climate of the Earth, Tellus Series A 21
611-619 (1969).
[2] Sellers W D, A global climatic model based on the energy balance of the Earth-atmosphere
system, Journal of Applied Meteorology 8 392-400 (1969).
[3] Kirschivink J L, Late proterozoic low-latitude global glaciation: the snowball earth, in The
Proterozoic Biosphere (J.W. Schopf and C. Klein, eds), pp. 51-52 Cambridge University Press,
Cambridge (1992).
[4] Hoffman P F, Kaufman A J, Halverson G P and Schrag D P, A neoproterozoic snowball
Earth, Science 281 1342-1346 (1998).
[5] Kennedy M J, Runnegar B, Prave A R, Hoffmann K H and Authur M A, Two or four
neoproterozoic glaciations, Geology 26 1059-1063 (1998).
[6] Hoffman P F and Schrag D P, The snowball Earth hypothesis: testing the limits of global
change, Terra Nova 14 129.155 (2002).
[7] Kaper H and Engler H, Mathematics and climate. Publisher: Society for Industrial and
Applied Mathematics, ISBN 978-1-611972-60-4 (2013).
[8] Flanner M G, Shell K M, Barlage M, Perovich D K and Tschudi M A, Radiative
forcing and albedo feedback from the northern hemisphere cryosphere between 1979 and 2008,
Nature Geoscience 4 151-155 (2011).
[9] Davy, R., Esau, I., Chernokulsky, A., Outten, S., and Zilitinkevich, S., Diurnal asym-
metry to the observed global warming, International Journal of Climatology (2016).
[10] Kasting J F, Whitmire D P and Reynolds R T, Habitable Zones around Main Sequence
Stars, ICARUS 101 108-128 (1993).
[11] Lyapunov A M, The general problem of the stability of motion, International Journal of
Control 53 531-534 (1992).
[12] Thompson J M T and Stewart H B, Nonlinear dynamics and chaos, Publisher: John Wiley
& Sons, ISBN 0-471-87684-4 (2001).
[13] Cvitanovi´c P, Artuso R, Mainieri R, Tanner G and Vattay , Chaos: Classical and
Quantum, ChaosBook.org (Niels Bohr Institute, Copenhagen) (2012).
[14] Zakrzhevsky M, KloKov A, Yevstignejev V and Shilvan E, Complete bifurcation anal-
ysis of driven damped pendulum systems, Estonian Journal of Engineering 17 76-87 (2011).
[15] Zevin A A, Existence and stability of forced oscillations in non-linear systems with one degree
of freedom, Int. J. Non-Linear Mechanics 30 205-221 (1995).
[16] Ott E, Grebogi C and Yorke J A, Controlling Chaos, Phys. Rev. Lett. 64 1196-1199 (1990).
[17] Shinbrot T, Grebogi C, Ott E and Yorke J A, Using small perturbations to control
chaos, Nature 363 411-417 (1993).
[18] Williams D M and Kasting J F, Habitable planets with high obliquities, ICARUS 129
254-267 (1997).
[19] Spiegel D S, Menou K and Scharf C A, Habitable climates: the influence of obliquity,
Astrophysical Journal 691 596-610 (2009).
[20] Dressing C D, Spiegel D S, Scharf C A, Menou K and Raymond S N, Habitable climates:
the influence of eccentricity, Astrophysical Journal 721 1295-1307 (2010).
[21] Schwartz S E, Heat capacity, time constant, and sensitivity of Earth's climate system, Journal
of Geophysical Research: Atmospheres 112 (D24) (2007).
[22] Schwartz S E, Determination of Earth's transient and equilibrium climate sensitivities from
observations over the twentieth century: strong dependence on assumed forcing, Surveys in
geophysics 33 (3-4), 745-777 (2012).
p-6
|
0903.1119 | 1 | 0903 | 2009-03-05T21:45:02 | Extreme Sensitivity of the YORP Effect to Small-Scale Topography | [
"astro-ph.EP"
] | Radiation recoil (YORP) torques are shown to be extremely sensitive to small-scale surface topography. Starting from simulated objects representative of the near-Earth object population, random realizations of three types of small-scale topography are added: Gaussian surface fluctuations, craters, and boulders. For each, the resulting expected relative errors in the spin and obliquity components of the YORP torque are computed. Gaussian power produces errors of order 100% if observations constrain the surface to a spherical harmonic order l < 10. A single crater with diameter roughly half the object's mean radius, placed at random locations, results in errors of several tens of percent. Boulders create torque errors roughly 3 times larger than do craters of the same diameter. A single boulder comparable to Yoshinodai on 25143 Itokawa, moved by as little as twice its own diameter, can alter the magnitude of the torque by factors of several, and change the sign of its spin component at all obliquities. A YORP torque prediction derived from groundbased data can be expected to be in error by of order 100% due to unresolved topography. Small surface changes caused by slow spin-up or spin-down may have significant stochastic effects on the spin evolution of small bodies. For rotation periods between roughly 2 and 10 hours, these unpredictable changes may reverse the sign of the YORP torque. Objects in this spin regime may random-walk up and down in spin rate before the rubble-pile limit is exceeded and fissioning or loss of surface objects occurs. Similar behavior may be expected at rotation rates approaching the limiting values for tensile-strength dominated objects. | astro-ph.EP | astro-ph | Extreme Sensitivity of the YORP Effect to Small-Scale Topography
Astrophysical Institute, Department of Physics and Astronomy, 251B Clippinger Research Laboratories, Ohio University, Athens, OH 45701, USA
Thomas S. Statler
9
0
0
2
r
a
M
5
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
1
1
1
.
3
0
9
0
:
v
i
X
r
a
Abstract
Radiation recoil (YORP) torques are shown to be extremely sensitive to small-scale surface topography, using nu-
merical simulations. Starting from a set of "base objects" representative of the near-Earth object population, random
realizations of three types of small-scale topography are added: Gaussian surface fluctuations, craters, and boulders.
For each, the expected relative errors in the spin and obliquity components of the YORP torque caused by the observa-
tionally unresolved small-scale topography are computed. Gaussian power, at angular scales below an observational
limit, produces expected errors of order 100% if observations constrain the surface to a spherical harmonic order
l . 10. For errors under 10%, the surface must be constrained to at least l = 20. A single crater with diameter roughly
half the object's mean radius, placed at random locations, results in expected errors of several tens of percent. The
errors scale with crater diameter D as D2 for D > 0.3 and as D3 for D < 0.3 mean radii. Objects that are identical
except for the location of a single large crater can differ by factors of several in YORP torque, while being photometri-
cally indistinguishable at the level of hundredths of a magnitude. Boulders placed randomly on identical base objects
create torque errors roughly 3 times larger than do craters of the same diameter, with errors scaling as the square of
the boulder diameter. A single boulder comparable to Yoshinodai on 25143 Itokawa, moved by as little as twice its
own diameter, can alter the magnitude of the torque by factors of several, and change the sign of its spin component at
all obliquities. Most of the total torque error produced by multiple unresolved craters is contributed by the handful of
largest craters; but both large and small boulders contribute comparably to the total boulder-induced error. A YORP
torque prediction derived from groundbased data can be expected to be in error by of order 100% due to unresolved
topography. Small surface changes caused by slow spin-up or spin-down may have significant stochastic effects on the
spin evolution of small bodies. For rotation periods between roughly 2 and 10 hours, these unpredictable changes may
reverse the sign of the YORP torque. Objects in this spin regime may random-walk up and down in spin rate before
the rubble-pile limit is exceeded and fissioning or loss of surface objects occurs. Similar behavior may be expected at
rotation rates approaching the limiting values for tensile-strength dominated objects.
Key words: ASTEROIDS, DYNAMICS, ASTEROIDS, ROTATION, ASTEROIDS, SURFACES, NEAR-EARTH
OBJECTS
1. Introduction
Radiation-recoil effects are now understood to be major drivers in the evolution of small bodies in the Solar
System. The secular drift in orbital elements produced by the Yarkovsky effect is a dominant mechanism in the spread
of asteroid families (Bottke et al., 2001) and in the replenishment of the near Earth object (NEO) population from the
inner main belt (Bottke et al., 2002). Torques produced by the YORP effect can compete with, and even dominate,
collisional and tidal torques on main belt asteroids and NEOs smaller than a few km in diameter (Rubincam, 2000;
Bottke et al., 2006, and references therein).
Recent observational efforts have achieved direct detection of the influence of the Yarkovsky and YORP mech-
anisms. Radar observations of 6489 Golevka show non-gravitational accelerations consistent with Yarkovsky pre-
dictions (Chesley et al., 2003). Photometric observations of 1862 Apollo (Kaasalainen et al., 2007) and 1620 Ge-
ographos ( Durech et al., 2008a), and photometric and radar observations of (54509) 2000 PH5 (subsequently named
Email addresses: [email protected] (Thomas S. Statler)
Preprint submitted to Icarus
May 25, 2018
YORP) (Lowry et al., 2007; Taylor et al., 2007) indicate accelerations in spin rate consistent with YORP models and
not attributable to tidal torques. However, the YORP models contain substantial systematic uncertainties, because the
objects' shapes and surface properties are not sufficiently constrained by observations. In the case of 1862 Apollo,
models computed by Kaasalainen (2004) with slightly different rotation pole positions and shapes span a range of
∼ ±25% around the measured acceleration. For 54509 YORP, the models computed by Taylor et al. (2007), using a
range of shapes and surface roughnesses, span a factor of ∼ 3 in predicted acceleration; and, in fact, they all systemat-
ically overestimate the observed effect, by factors of 2 to 7.1 Considering the extensive observational effort devoted to
these objects, the ambiguity in the results raises considerable concern as to whether YORP predictions derived from
groundbased light curve and/or radar data can ever be sufficiently precise to permit quantitative tests of the theory.
This concern is only deepened by the history of YORP calculations made for 25143 Itokawa before and after the
arrival of the Hayabusa spacecraft. Using pre-encounter shape models derived from radar observations (Ostro et al.,
2004) and photometric light curves (Kaasalainen et al., 2003), Vokrouhlick´y et al. (2004) predicted a rotational accel-
eration of roughly half a part in 104 per year, with an estimated uncertainty of ±30%. Scheeres et al. (2007) subse-
quently recomputed the YORP prediction, first from a similar pre-encounter model (Gaskell et al., 2006), obtaining a
result consistent with that of Vokrouhlick´y et al. (2004), and then from a sequence of shape models of progressively
higher resolution based on in situ Hayabusa data. The first post-encounter shape model was found to imply a rota-
tional deceleration, meaning that pre-encounter models had been unable to correctly predict even the sign of the effect.
Subsequent models incorporating the later Hayabusa data resulted in YORP accelerations that did not converge as the
resolution increased. Scheeres et al. (2007) conclude that, even at sub-meter resolution, the essential geometry of the
interaction between the surface and the incoming and outgoing radiation is not adequately represented. In the past year
the situation has become even murkier. Scheeres and Gaskell (2008) show that the torque on 25143 Itokawa is very
sensitive to the position of the center of mass; and the predicted acceleration has still not been detected ( Durech et al.,
2008b).
The foregoing stories motivate this paper, which asks whether extreme sensitivity to small-scale surface topogra-
phy is a characteristic of the YORP effect itself, or is limited to a comparatively small number of unfortunately shaped
objects. This question has pragmatic as well as predictive consequences. First, dynamically significant small scale
topography may be observationally unresolvable. I will show below that, indeed, neglecting topographic features,
such as moderate sized craters, that are indistinguishable from the ground can lead to errors in YORP predictions of
order unity. As a result, shape models derived from groundbased data are unlikely ever to yield precise predictions
of YORP torques. Second, dynamically significant small scale topography may, in some situations, be unpredictably
changeable. I will demonstrate that a small displacement of a single large boulder can, in the right circumstances,
completely reverse the sign of the YORP torque. This implies that a YORP cycle (Rubincam, 2000) can be inter-
rupted in a major way by a comparatively minor event, and suggests that the spin evolution of small NEOs may be
more stochastic than previously recognized.
The theoretical foundations and basic behavior of the YORP effect have been analyzed extensively by analytic,
semi-analytic, and numerical means (e.g., Scheeres, 2007b; Nesvorn´y and Vokrouhlick´y, 2007, 2008; Breiter and Michalska,
2008; Micheli and Paolicchi, 2008; Mysen, 2008b,a). These works have elucidated how asymmetries in the illumi-
nated and radiating surface couple to produce the now familiar behavior of the orbit-averaged torque components as a
function of obliquity (Vokrouhlick´y and Capek, 2002), and have provided strategies for computing torques for realis-
tic asteroid models. This paper adopts a purely numerical treatment, analyzing a large number of simulated objects to
yield statistically meaningful results in the nonlinear regime where shadowing is important.
The remainder of the paper is laid out as follows. In Section 2, I describe the calculation of YORP torques for
simulated asteroids. I define a set of "base objects" on which the effect of small-scale topography will be tested, and
describe how topography is added to the objects, in the form of Gaussian surface fluctuations, craters, or boulders. In
Section 3, I demonstrate that small scale topography, at a level difficult or impossible to constrain from groundbased
observations, can typically be expected to alter YORP torques by tens of percent or more. In particular, the magnitude,
and even the sign, of the torque can hinge on the location of a single large crater or boulder. In Section 4, I discuss the
implications of these results, both for accurate predictions of the YORP effect and for the YORP-driven evolution of
small NEOs.
1Scaling down the models by these factors was necessary to achieve the excellent fit in Fig. 2 of Lowry et al. (2007), as explained in the caption
to that figure.
2
2. Method
2.1. Calculation of Torques
Radiation interacting with the surface of a body can impart a torque through the momentum it carries. It is concep-
tually helpful to split the interaction into three parts: (1) the momentum deposited on the surface by arriving photons;
(2) the recoil imparted to the surface by departing reflected photons; and (3) the recoil imparted by departing reradiated
photons. In this paper I focus specifically on the third contribution, the reradiation torque. The first contribution can
be shown to cancel identically when averaged around an orbit (Nesvorn´y and Vokrouhlick´y, 2008). The second is pro-
portional to the albedo, and becomes insignificant for a sufficiently dark surface; and, as Nesvorn´y and Vokrouhlick´y
(2007) point out, if the reflected and reradiated intensities are both assumed isotropic, then the torque contributions
are parallel.
I compute the torques on simulated objects using the TACO2 code, developed by the author and students at Ohio
University. The object's surface is represented using a standard triangular tiling (Lagerros, 1996). For all results
presented here, 6,394 vertices and 12,784 tiles are used, for an average angular resolution of approximately 7.5◦ (3
tiles wide). For comparison, typical shape models derived from asteroid light curve observations have 1,000 to 2,000
vertices and 2,000 to 4,000 tiles (e.g., Durech et al., 2007). Models obtained from radar observations and currently
archived in the Planetary Data System (PDS) have typically around 4,000 vertices and 8,000 tiles. Using groundbased
data alone, currently only 4 objects (1620 Geographos, 25143 Itokawa, 4179 Toutatis, and 52760) have been modeled
at resolutions equal to or finer than that used in this paper.
A horizon map (the maximum elevation ψh of all visible parts of the surface as a function of azimuth η, measured
relative to local west) is computed at the centroid of each tile when the surface is initially defined. Shadowing is
handled by requiring that the Sun not be blocked by other parts of the surface for the tile to be illuminated. An
illuminated tile absorbs a flux Fabs given by
Fabs = Fµ0(cid:2)1 − rhem(µ0)(cid:3) ,
(1)
where F is the incident flux, µ0 is the cosine of the angle between the incident flux and the surface normal, and
rhem is the directional-hemispheric reflectance or hemispheric albedo. TACO computes rhem using eq. (42) of Hapke
(2002), assuming a single Henyey-Greenstein phase function. For definiteness I adopt the mean Hapke parameters for
S-class asteroids given by Helfenstein and Veverka (1989): a single-scattering albedo w = 0.23, and a phase function
expanded to 6th order in Legendre polynomials with a width parameter ξ = −0.35. The correction for macroscopic
surface roughness from Hapke (1984) is applied with a mean slope parameter ¯θ = 20◦; this has no bearing on the
absorbed flux but is relevant to the computation of light curves (section 3). These reflectance parameters are assumed
to be constant over the surface. The opposition effect (Hapke, 2002) is ignored, as is the illumination of other parts of
the surface by the reflected flux.
I neglect the effects of thermal conductivity, so that the reradiated flux exactly balances the absorbed flux at all
times. Each tile is assumed to radiate isotropically; if there is no blockage by the local horizon, the outgoing photons
uniformly fill the upward-facing hemisphere, and the recoil force on the tile is given by
frec = 2πZ π/2
0
FabsA
πc
cos2 θ sin θdθ =
2FabsA
3c
,
(2)
where A is the tile area, c is the speed of light, and θ is the polar angle from the surface normal. In this case the
recoil force is directed inward, normal to the surface. If, instead, part of the sky is blocked by the local horizon, some
reradiated photons from the tile will strike other parts of the object, and the tile will intercept some photons reradiated
by these other parts. Photons traded between different parts of the surface cannot exert a torque on the body until
they are radiated into the clear sky. As a result, both the magnitude and direction of the recoil force will be changed.
I compute an approximate correction for this effect, assuming that traded photons have no net effect on the energy
absorbed by each tile. For the moment, suppose that the tile's local horizon were uniformly elevated an angle ψh
2"Thermophysical Asteroid Code, Obviously"
3
above the horizontal. Then for the same radiated flux, the intensity would be increased by a factor sec2 ψh, and the
magnitude of the recoil force would be given by
frec = 2πZ π/2−ψh
0
FabsA
πc cos2 ψh
cos2 θ sin θdθ =
2FabsA
3c
1 +
sin2 ψh
1 + sin ψh! .
(3)
The force would be increased, relative to the flat-horizon case, by a factor 1 + sin2 ψh/(1 + sin ψh) because the elevated
horizon causes the outgoing radiation to be "beamed" into a narrower solid angle. In practice, of course, the elevation
of the horizon is not uniform in azimuth. Therefore I replace ψh with the mean elevation, given by
¯ψh =
1
Nh
Nh−1
Xi=0
ψh(ηi),
(4)
where ηi is the azimuth of the ith point (out of Nh = 50) in the horizon map. The magnitude of the recoil force is then
taken to be amplified by a factor 1 + sin2 ¯ψh/(1 + sin ¯ψh) over the flat-horizon case. The direction of ~frec is taken to be
antiparallel to the clear-sky normal Nsky, which is computed from a first-order Fourier expansion of the horizon map:
Nsky ≡
~nsky
nsky
,
(5)
(6)
,
where
~nsky = Nsurf − ǫn tan
2
Nh
Nh−1
Xi=0
ψh(ηi) cos ηi
− ǫw tan
2
Nh
Nh−1
Xi=0
ψh(ηi) sin ηi
Nsurf is the surface normal, and ǫn and ǫw are the unit vectors in the local north and west directions, respectively.
The total instantaneous torque is then simply the cross product ~r × ~frec, where ~r is the vector from the center of
mass to the tile centroid, summed over tiles. The torque is averaged over the spin and orbital phase, assuming the
periods are incommensurable. For computational simplicity, the orbit is taken to be circular. For orbits with finite
eccentricity e, the averaged torque can be simply scaled by a factor (1 − e2)−1/2 (Scheeres, 2007b).
2.2. Base Objects
I define a set of base objects, on which to measure the effects of smaller scale topography, using the "Gaussian
random sphere" formalism of Muinonen and Lagerros (1998). In this approach the distance to the surface from the
origin is given by
where (θ, φ) are the usual spherical-coordinate angles and the function s(θ, φ) is defined by an expansion in spherical
harmonics:
slmYlm(θ, φ).
(8)
In this formulation, the mean radius of the body ¯r ≡ 1. Since s(θ, φ) is real, the coefficients slm are constrained by the
relation
r(θ, φ) = exp"s(θ, φ) −
β2
2 # ,
∞
l
s(θ, φ) =
Xl=0
Xm=−l
sl,−m = (−1)ms∗
lm.
(7)
(9)
For nonnegative m, each slm is a Gaussian random variable having zero expectation value and a variance given by
Var[ℜ(slm)] = (1 + δm0)
Var[ℑ(slm)] = (1 − δm0)
2π
2l + 1
2π
2l + 1
Cl;
Cl.
(10)
In eq. (10), δi j is the Kronecker delta symbol; Cl represents the coefficients of the Legendre expansion of the log-
radius covariance function, which also define the quantity β ≡ (Pl Cl)1/2 in eq. (7). Muinonen and Lagerros (1998)
4
Figure 1: Coefficients of the Legendre expansion of the log-radius covariance function. Squares: as tabulated by Muinonen and Lagerros (1998).
Diamonds and dotted line: from the adopted fit, Eq. (11).
Figure 2: Crater profile according to Eq. (12), with depth to diameter ratio q = 0.2 and rim width δ = D/20.
estimate the Cl coefficients for l ≤ 10 from shapes fitted to light curve observations of 14 objects. For this work I
adopt an analytic fit to the results for their small-object subsample (4th column of their Table 3), given by
Cl = 1.2
(l2 + 0.26)2
(l8 + 90.0)1.06 .
(11)
Fig. 1 shows the coefficients from Muinonen and Lagerros (1998) and the fit. This fit will be used below to extrapolate
the covariance function out as far as l = 40. For l ≫ 1, Eqs. (10) and (11) imply Cl ∼ l−4.48, so that the RMS amplitudes
slm decline as l−2.74.
Each base object is a Gaussian random sphere with the expansion in Eq. (8) truncated at a maximum order lbase.
I adopt values of lbase = 4, 5, 7, 9, 12, 15, 20, and, for each value, compute 6 different base objects with independent
random realizations of the slm coefficients.
2.3. Addition of Small-Scale Topography
The base objects are intended to represent asteroids whose shapes are constrained by observations down to some
minimum angular scale. Different types of small-scale topographic structure can be added to the base objects. In
5
Figure 3: Typical variation in YORP torques caused by Gaussian topographic power added at smaller scales to the base objects. (a) Base object
with lbase = 9, topography added to lmax = 18. Top- and bottom-most curves correspond to the objects shown in Fig. 4. (b) Base object with
lbase = 20, topography added to lmax = 40. Top panels show the spin component, and bottom panels the obliquity component, of the torque,
normalized by the moment of inertia about the short (spin) axis, plotted against obliquity. Individual curves correspond to ten different random
realizations of the added topography.
Section 3 I measure the separate effects on YORP of each of the following:
1. Gaussian power at smaller scales. Here the expansion in Eq. (8) is simply extended from lbase to some higher
maximum order lmax. The slm coefficients are identical to those for the base object for l ≤ lbase, and chosen
randomly according to Eq. (10) for lbase < l ≤ lmax.
2. Craters. One or more circular craters are placed randomly on the surface. The crater depth z as a function of
radius r from its center is given by
z(r) = qD 1 − 4
r2
D2! ×( 1, if r ≤ D/2;
exph− (r−D/2)2
2δ2
i , if D/2 < r ≤ D/2 + 3δ.
(12)
In Eq. (12), D is the crater diameter, q is the depth-to-diameter ratio, and δ is the width of the raised rim. I
adopt q = 0.2 and δ = D/20, based on NEAR-Shoemaker observations of 253 Mathilde (Thomas et al., 1999)
and 433 Eros (Veverka et al., 2001; Thomas et al., 2002) and laboratory impact experiments (Nakamura, 2002;
Housen and Holsapple, 2003). The resulting profile is shown in Fig. 2. The vertical displacement z is in the
direction of the mean surface normal of the area the crater will occupy. Prexisting topography is first flattened to
the level of the mean surface and then overwritten by the crater. Multiple craters are added sequentially so that
overlapping features are reasonably realistic.
3. Boulders. One or more boulders can also be placed on the surface. As a crude representation of a boulder, the tile
vertices falling within a given radius of the boulder center are raised, creating an approximately round, mesa-like
feature with jagged edges. As with craters, the existing topography is first flattened so that the boulder's upper
surface is featureless.
For each object, the center of mass and inertia tensor are calculated after the small-scale topography is added, assuming
a homogeneous density ρ. The coordinate system is shifted and rotated to align the origin with the center of mass
and the coordinate axes with the principal axes of the body. All objects are assumed to be rotating about the shortest
principal axis.
3. Results
3.1. Gaussian Power at Smaller Scales
Any set of observations will be able to constrain the topographic power down to some minimum length scale or
maximum harmonic order. As a straightforward indicator of the effect of neglected higher orders, I add power to each
6
Figure 4: Identical views of the two lbase = 9 objects from Fig. 3a with the (left) smallest and (right) largest torques. Line of sight is in the plane
containing the short and middle axes of the body, 45◦ from each, and the illumination is from the right at 45◦ phase angle.
base object to a maximum order lmax = 2lbase, and measure the error in torque incurred by ignoring topography down
to half the observationally constrained scale.
Fig. 3 shows typical results for two of the 42 base objects, one with lbase = 9 and one with lbase = 20, corresponding
to minimum angular scales of 40◦ and 18◦, respectively. The curves show the averaged torque components affecting
the spin rate (τs) and the obliquity (τǫ), normalized by the principal moment of inertia about the short axis Is, plotted
against obliquity.3 Different curves indicate the 10 random realizations of the added topography. Clearly, knowing
the topography only to order lbase = 9 (Fig. 3a) is not sufficient to determine the torque to even order-unity accuracy.
The range in τs spans nearly a factor of 3 at obliquity ǫ = 0◦, and more than a factor of 5 at ǫ = 90◦. The location
of the "Slivan states" (Slivan, 2002), which require τs = 0, varies over 15◦; and the obliquity torque τǫ in the vicinity
of the Slivan states varies over a factor of 5. In contrast, the lbase = 20 objects (Fig. 3b), to which power has been
added to lmax = 40, have largely concordant torques. Evidently, at sufficiently small angular scales, the neglect of
Gaussian power at still smaller scales is not too serious, as long as it follows a simple extrapolation of the covariance
coefficients.
As a measure of how wrong a YORP prediction is likely to be due to neglect of the smaller scale topography, I
compute, at each value of obliquity, the standard deviation in each torque component over the 10 trials, then average
over obliquity; this quantity is normalized to the mean absolute value of that component, also averaged over obliquity.
The result is an obliquity-independent measure of the expected relative error. For the lbase = 9 case in Fig. 3a, the
expected relative errors are 38% in τs and 32% in τǫ. Keep in mind that these are 1σ errors, implying 95% confidence
bands (for a Gaussian distribution) with full widths of 152% and 128%, consistent with the order-unity variation seen
in the figure. For the lbase = 20 case, the relative 1σ errors are 7% and 3% in τs and τǫ respectively. I will use this
same approach to quantify the effects of craters and boulders below.
The most remarkable thing about the results in Fig. 3a is that the individual objects, despite differing by factors
of a few in their YORP torques, appear nearly identical. Fig. 4 shows a representative view of the two objects in
Fig. 3a having the most discordant torques. To assess whether these objects would be distinguishable photometrically,
I compute optical light curves over a grid of 40 rotation pole directions and 10 phase angles from 0◦ to 90◦. I find that
the RMS photometric difference between these two objects is < 0.05 magnitudes at all phase angles. The maximum
difference is < 0.1 mag at phase angles ≤ 50◦, and is > 0.2 mag only at phase angles of 75◦ or higher, making it
unlikely that these objects would be distinguished by groundbased observations.
In Fig. 5, I show the relative 1-sigma errors for all of the base objects. The scatter at each lbase reflects the base
objects' different YORP susceptibilities. One can easily see that order-unity errors in the torque should be expected for
lbase ≤ 10. To be reasonably confident of errors under 10%, the surface must be known to at least lbase = 20. Owing to
3Numerical results for normalized torques are given in units of 1.52 × 10−15(ρ/2 g cm−3)−1(¯r/1 km)−2(a/1 AU)−2(1 − e2)−1/2 s−2, where ρ and
¯r are the constant density and mean radius of the body, and a and e are the orbital semimajor axis and eccentricity.
7
Figure 5: Expected relative 1σ error in the YORP torque caused by neglecting Gaussian topographic power at lbase < l ≤ 2lbase. Top: error in the
spin component; bottom: error in the obliquity component. The relative error is the standard deviation in each component over 10 random trials,
averaged over obliquity, and normalized to the mean absolute value of that component. Multiple symbols correspond to 6 different base objects at
each lbase.
the steepness of the extrapolated covariance function, most of the change in the torque is actually produced by orders
close to lbase. Fig. 6 shows the relative errors for the lbase = 12 objects with power added to lmax = 13, 14, 16, 18, 24,
and 36. There is no discernible increase in the error for lmax > 18, and the single order l = 13 seems to be responsible
for roughly half of the asymptotic value.
3.2. Craters
Actual small-scale topography is likely to be non-Gaussian. Impact craters are one example of Poisson-distributed
features not well described by a truncated spherical-harmonic expansion. The crater size distribution has been esti-
mated for 951 Gaspra (Chapman et al., 1996b; Greenberg et al., 1994), 243 Ida (Belton et al., 1994; Chapman et al.,
1996a), 253 Mathilde (Chapman et al., 1999), and 433 Eros (Chapman et al., 2002). To lowest order, the cumula-
tive distribution of craters with diameters larger than D, N(D), is reasonably well approximated by a power law,
N(D) ∼ D−2 (O'Brien et al., 2006). Though there are significant deviations in the logarithmic slope for different
objects, the known distributions all describe a steeply diminishing number of large craters. The largest craters have
diameters ranging from 0.35 to 1.2 times the bodies' mean radii (Thomas et al., 1999).
I return to the original base objects, removing the small-scale Gaussian power added in Section 3.1, in order
to measure separately the effect of craters on the YORP torque components. I start by randomly placing a single,
modest-sized crater of diameter D = 0.3 (in units of the body mean radius) on the base objects. Fig. 7 shows the
resulting relative 1σ errors, calculated as in Sec. 3.1 with ten random crater placements for each object. A single
crater can clearly alter the spin and obliquity torques by tens of percent, even though the crater itself, including the
raised rim, occupies only ∼ 0.5% of the total surface. The effect increases with lbase at values lbase . 9. A similar,
but less pronounced, increase is also seen in the Gaussian case (Fig. 5). The single crater actually produces less error
8
Figure 6: Expected relative 1σ error, as in Fig. 5, for the lbase = 12 base objects with Gaussian power added to lmax = 13, 14, 16, 18, 24, 36. There
is no strong dependence on lmax for lmax > 18, indicating that orders just above the cutoff lbase are most important.
for lbase . 9 than does small-scale Gaussian power added to the same base object. But the effect of the crater for
lbase > 10 is larger than the Gaussian case and, to within the statistical errors, independent of the base order. This may
be due either to the declining amplitudes Cl at higher l or to the fact that l = 9 corresponds roughly to the scale of the
crater itself. Evidently the crater couples to other mesoscale surface features of roughly its own size more effectively
than to the large-scale structure at smaller l.
Fig. 8a shows the normalized torque as a function of obliquity for ten random placements of single craters on
the same lbase = 20 base object used in Fig. 3b. The effect of the crater is about twice that of the added Gaussian
topography: the relative errors in τs and τǫ are 14% and 6%, respectively, which also represent median results over all
base objects. Fig. 8b shows the effect of doubling the crater size to D = 0.6 on the same base object. The relative errors
increase to 56% and 14%. Comparing the extreme examples (the highest and lowest curves), the spin components
differ by a factor of 4.4 at ǫ = 90◦ and actually have opposite signs at ǫ = 0◦. Fig. 9 shows these two most discordant
objects, which turn out to be cratered at almost diametrically opposite positions.
Cratering a base object moves the center of mass and slightly reorients the principal axes. To determine whether
these adjustments contribute to the change in torque, I recompute the torques for 60 of the cratered and bouldered
(see Section 3.3) objects, without shifting to the new center of mass or reorienting to the new principal axes. For
either craters or boulders, the results differ, on average, by only a few percent, indicating that it is really the surface
topography that is responsible for the large expected torque errors.
The objects, even those with single large (D = 0.6) craters, are so photometrically similar that it is unlikely that
they would be distinguishable in current light curve programs. To assess this possibility, I take all 60 of the lbase = 20
cratered objects and compute light curves over the same grid of 40 orientations and 10 phase angles used in section
3.1. I then calculate the probability that two objects, differing only in crater location, would be observed at a given
moment to have a magnitude difference > ∆. The result, as a function of phase angle, is shown in Fig. 10. Note
9
Figure 7: Expected relative 1σ standard deviation in the YORP torque, as in Fig. 5, caused by a single randomly placed crater with diameter
D = 0.3 added to different base objects.
that, unless substantial data are obtained at phase angles > 70◦, one expects that no more than 1% of the total data
will differ by more than 0.05 mag, and no more than 0.01% by more than 0.1 mag. In practice, observations at phase
angles above 45◦ are rare in light curve programs.
Fig. 11 shows the scaling of the relative torque errors with crater size. The diamonds and plus signs show results
for single-cratered objects with lbase = 12 and 20, respectively. One might naively expect the torque error to scale with
the crater area (∼ D2); and Fig. 11 shows that this is not a bad description for D > 0.3. However, at smaller diameters
the scaling is much steeper, with the relative error roughly ∝ D3. This nonlinear response may be attributable to
shadowing, although the same steepening does not occur for boulders (Section 3.3 below). The squares and crosses
in Fig. 11 (slightly offset for clarity) indicate the results for objects with four D = 0.3 craters each. Since the changes
in torque produced by each added crater are random and uncorrelated, one should expect the total change to scale as
the square root of the number of craters. This expectation is confirmed by the factor 2 increase in the relative errors
over the single-crater case. The implications of these scaling laws for the expected accuracy of YORP predictions on
realsitically cratered objects are discussed in section 4 below.
3.3. Boulders
Boulders are a second example of Poisson-distributed topography. The surface of 25143 Itokawa has been found
to be dominated by boulders, with a cumulative size distribution well represented by N(D) ∝ D−3 over a factor 10 in
diameter; the largest boulder, Yoshinodai, is fully one-tenth the size of its parent body (Saito et al., 2006). For 433
Eros, Chapman et al. (2002) obtain a result consistent with N(D) ∝ D−4. These size distributions are substantially
steeper than those for craters; thus one can expect the cumulative effect of moderate sized boulders to be more
relatively important than that of moderate sized craters.
I proceed with the same strategy to assess the influence of boulders on YORP torques. I add a single mesa-shaped
boulder, with diameter D = 0.3 (in units of the body mean radius) and height h = D/2, at random locations to the base
10
Figure 8: Typical variation in YORP torques caused by single craters of diameter (a) D = 0.3, (b) D = 0.6, randomly placed on the same base
object used in Fig. 3b. Top- and bottom-most curves correspond to the objects shown in Fig. 9.
Figure 9: Identical views, along the rotation axis, of the two objects having the (left) smallest and (right) largest torques in Fig. 8b.
objects. The base objects have mean longest diameters of approximately 3.1, so the size of these boulders relative
to their parent bodies approximately matches that of Yoshinodai. The height-to-diameter ratio corresponds to typical
values measured for boulders on 433 Eros (Thomas et al., 2002).
Fig. 12 shows the expected relative error in the torque components caused by the single large boulder. The
effect is quite dramatic: typical 1σ errors are 10% to 100% in the spin component, and 5% to 50% in the obliquity
component. This is roughly a factor 3 larger than the effect of craters of the same diameter, though the difference may
be attributable to the simple fact that boulders can be proportionally taller than craters are deep (here, by a factor of
2.5). As in the case of craters, there is a weak increase in the mean error for values of lbase < 10, but little difference
is seen at higher order. The scaling of the errors with boulder size is shown in Fig. 13. The boulders are all similarly
shaped, with h = D/2. The relative error scales quite cleanly as D2, i.e., proportional to the boulder surface area.
Examples of the torque components as a function of obliquity are shown in Fig. 14a, for 10 random D = 0.3
boulder placements on the same lbase = 20 object used in Figs. 3b and 8. This case again represents a typical amount
of topographically induced variation. Boulders have a tendency to shift the curves for τs/Is up and down while altering
the overall amplitude of those for τǫ /Is. Note that the sign of the spin torque can be flipped by the location of a single
boulder, even for this unexceptional object. A more extreme example is seen in Fig. 14b, which shows the same
curves for one of the other five lbase = 20 objects. Here the sign of the spin torque at all obliquities can be reversed by
the location of a single Yoshinodai-sized boulder. The two most discordant objects are shown in Fig. 15. The boulder
11
Figure 10: Probability contours for observing a photometric difference ∆ or larger, as a function of phase, between pairs of objects that differ only
in the location of a single crater of diameter D = 0.6. Results shown are averaged over the 6 base objects with lbase = 20.
location differs by barely twice its own diameter; yet the object on the left will inexorably spin up while the object on
the right will spin down.
4. Discussion
The above results indicate that sensitivity to small-scale topography is an inherent characteristic of the YORP
effect. Since YORP operates by virtue of the non-cancellation of many independent torque components that vary
across the surface, over the rotation of the body, and around the orbit, this conclusion is perhaps not surprising;
however, it does impose strong limits on the predictability of YORP torques, and has important consequences for the
spin evolution of small bodies.
If the surface is describable by a spherical harmonic expansion whose coefficients are Gaussian-distributed, the
topography must be known accurately at least to harmonic order l = 10 to allow a YORP prediction of even the correct
sign. To achieve 10% accuracy requires the surface to be known at least to l = 20, if the topography diminishes with
l in a manner consistent with a simple extrapolation of the Muinonen and Lagerros (1998) covariance coefficients
(Eq. [11]). Few objects can be modeled reliably at this resolution. Furthermore, actual small scale topography is not
Gaussian and not small in amplitude. The above results show that Poisson-distributed features such as craters and
boulders, which produce substantial shadowing and which are not well revealed by groundbased observations, can
have significant effects. A single large feature can alter YORP torques by tens of percent or more.
I found above that the relative torque error scales as σ ∼ Dβ, where β ≈ 2 for boulders and 2 . β . 3 for craters.
Using these numerical results and the observed size distributions, one can estimate the relative influence of large and
small craters and boulders. The observed cumulative counts are reasonably well represented by N(D) ∼ D−α, with
12
Figure 11: Expected relative error in torque, as in Fig. 7, caused by single craters randomly placed on base objects with lbase = 12 (plus signs) and
lbase = 20 (diamonds), as a function of crater diameter D. Crosses and squares (offset horizontally) show results for 4 randomly placed craters each
of diameter D = 0.3, on objects with lbase = 12, 20 respectively.
3 . α . 4 for boulders and 1 . α . 2 for craters.4 Assuming the effects of individual boulders or craters to be
uncorrelated, one expects in either case that the variance in torque produced by features with diameters between Dmin
and Dmax will scale according to
σ2 ∝ Z Dmax
Dmin
dN
dD
D2βdD ∝ D2β−α
max − D2β−α
min ,
(α , 2β).
(13)
For cratered objects, Eq. (13) implies that at least 75% of the total variance (87% of the expected σ) is produced
by the 4 biggest craters. Fig. 11 shows that the expected error from the random location of the single largest crater is
already approaching 100% at Dmax = 0.6, close to the observed maximum diameter on Eros but rather modest com-
pared to the values of 0.89 and 1.2 for Ida and Mathilde, respectively (Thomas et al., 1999). These very large craters
could possibly give rise to noticeable features in high phase-angle light curves; however, light curve inversion per-
forms poorly in recovering surface concavities because the inversion problem is non-unique (Kaasalainen et al., 1992;
Kaasalainen and Torppa, 2001). The convex, pre-encounter model of the Rosetta target 2867 Steins (Lamy et al.,
2008) resembles the in situ images,5 but by construction misses several large craters, including one with an apparent
diameter D ≈ 0.9. A non-convex inversion probably would not have revealed this crater, as the light curve data are
limited to phase angles < 42◦. Convex models, in general, may be useful for determining overall surface reflectance
and thermal properties, and may give accurate predictions for Yarkovsky accelerations. But a YORP prediction based
on a convex model is likely to be wrong by of order 100%.
4As before, diameters are implicitly in units of the body mean radius.
5http://www.esa.int/SPECIALS/Rosetta/SEMNMYO4KKF 0.html
13
Figure 12: Expected relative 1σ standard deviation in the YORP torque, as in Fig. 5, caused by a single randomly placed boulder with diameter
D = 0.3 added to different base objects.
Boulders, because of their steeper size distribution, behave differently: at most half of the total variance is pro-
duced by boulders larger than half the maximum size. In fact, if the distribution is as steep as on Eros (α = 4), then
the variance is formally divergent as D → 0. Of course, distributions dN/dD ∼ D−1 or steeper are unphysical in
this limit; but the scaling indicates that the YORP torque can be affected as much by the surface distribution of small
boulders as by that of large ones. The results of section 3 also show that single boulders may have significantly greater
effect on the torque than craters of the same size; a single Yoshinodai-sized object can easily alter the spin torque by
tens of percent, and in some cases change its sign. The clear implication is that the surface boulder distribution can be
critical in determining the actual consequences of YORP.
Thus, one should expect uncertainties of order unity in predictions of the YORP effect made from models derived
from remote observations, just from the unconstrained nature of the small-scale surface topography. The true situation
is actually even worse, since I have neglected other issues, such as variations in albedo, surface roughness, and internal
density, which probably have comparable effects. Indeed, Scheeres and Gaskell (2008) find that shifting the center of
mass of Itokawa by only 15 m can reverse the sign of the spin torque. At present, there have been 4 attempts to predict
the YORP acceleration from groundbased data and compare with direct measurements. In one case (1682 Apollo),
the correct magnitude and sign were obtained, with an uncertainty of ±25%; in a second (54509 YORP), models
gave the correct sign but the wrong magnitude by a factor of 2 to 7. In the third (25143 Itokawa), pre-Hayabusa
models evidently obtained the wrong sign (compared to later models), but the predicted acceleration has still not been
detected. And in the fourth (1620 Geographos), torques computed for different convex shape models bracketed the
observed value, but varied over a factor of 5. This situation is about what we should expect. Generally speaking, a
YORP prediction computed from a groundbased model will be accurate only to within factors of order unity, and will
have no better than an 80% chance of having the correct sign.
The sensitivity of reradiation torques to boulders also implies that there may be substantial stochasticity in the spin
14
Figure 13: Expected relative error in torque, as in Fig. 12, caused by single boulders, as a function of boulder diameter D. Symbols have the same
meaning as in Fig. 11.
evolution of small bodies under the influence of YORP. In the standard picture of the YORP cycle (Rubincam, 2000),
an object can gradually spin up, simultaneously evolving in obliquity, as determined by the now familiar "YORP
curves" (e.g., Figs. 3, 8, and 14). In most cases, spin-up continues only until reaching an orientation regime where
τs < 0. The object then spins down, asymptotically approaching the stable fixed point in obliquity. Objects with
substantial power in the higher YORP orders (Nesvorn´y and Vokrouhlick´y, 2007, 2008) can sometimes have stable
fixed points at obliquities where τs > 0; examples of this behavior are seen in Figs. 8b and 14b, along with cases where
τs > 0 at all obliquities. These objects would presumably spin up indefinitely. But a new possibility raised here is
that a seemingly minor topographic change, such as the movement of a boulder, can stochastically shift the evolution
onto a new set of YORP curves. One would expect such events to occur well before catastrophic fissioning or even
unbinding of surface material, as the local effective gravity vectors change with the accelerating spin. Calculations
for contact-binary configurations (Scheeres, 2007a) and numerical experiments on rubble piles (Walsh et al., 2008;
Harris et al., 2009) confirm that surface changes occur at rotation rates slower, by factors of several, than the nominal
centrifugal limit for self-gravitating spheres. The latter corresponds to the observed minimum rotation periods of large
asteroids, at approximately 2 hours. Thus a broad range of spin periods, from 2 to perhaps 10 hours, may correspond
to a "scattering zone", in which small topographical changes stochastically reverse the YORP torques. An object
being spun up by YORP may scatter multiple times through motion or loss of surface boulders, random-walking up
and down in spin rate, before finally crossing the actual spin limit and either fissioning or shedding all loose objects.
It is extremely interesting that the observed distribution of NEOs in the diameter-rotation period plane (e.g., Fig. 1 of
Holsapple, 2007, Statler et al. 2009, in preparation) shows the sharp cutoff at 2 hours becoming notably indistinct at
diameters . 1 km, where YORP is expected to become dominant. Similarly, small, fast-rotating objects approaching
the higher spin limit imposed by tensile strength (Holsapple, 2007) may wander up and down in spin rate, driven not
by catastrophic fragmentation, but merely by a subtle rearrangement of the surface that reverses the YORP effect.
15
Figure 14: Typical variation in YORP torques caused by single boulders of diameter D = 0.3, randomly placed on (a) the same base object used
in Figs. 3b and 8, and (b) a second object with lbase = 20, showing more extreme variation. Top- and bottom-most curves correspond to the objects
shown in Fig. 15. Note in (b) the possibility of reversing the sign of the spin torque at all obliquities simply by repositioning one boulder.
Figure 15: Identical views of the two objects from Fig. 14b with the (left) most positive and (right) most negative spin torques. Line of sight is in
the plane containing the short and middle axes of the body, 45◦ from each, and the illumination is from the right at 30◦ phase angle. Moving the
boulder by only twice its own diameter reverses the sign of the spin torque.
The author is grateful to David Riethmiller, Desiree Cotto-Figueroa, and Kyle Uckert for their assistance in the
development of TACO; to Nalin Samarasinha, Dan Scheeres, Mangala Sharma, and Desiree Cotto-Figueroa for ex-
tensive comments on the manuscript; and to the referees, David Vokrouhlick´y and David Rubincam, for important
suggestions that improved the paper. This work has made use of NASA's Planetary Data System and Astrophysics
Data System Bibliographic Services.
References
Belton, M. J. S., Chapman, C. R., Veverka, J., Klaasen, K. P., Harch, A., Greeley, R., Greenberg, R., Head, III, J. W., McEwen, A., Morrison, D.,
Thomas, P. C., Davies, M. E., Carr, M. H., Neukum, G., Fanale, F. P., Davis, D. R., Anger, C., Gierasch, P. J., Ingersoll, A. P., Pilcher, C. B.,
1994. First Images of Asteroid 243 Ida. Science 265, 1543–1547.
Bottke, W. F., Morbidelli, A., Jedicke, R., Petit, J.-M., Levison, H. F., Michel, P., Metcalfe, T. S., 2002. Debiased Orbital and Absolute Magnitude
Distribution of the Near-Earth Objects. Icarus 156, 399–433.
Bottke, W. F., Vokrouhlick´y, D., Broz, M., Nesvorn´y, D., Morbidelli, A., 2001. Dynamical Spreading of Asteroid Families by the Yarkovsky Effect.
Science 294, 1693–1696.
Bottke, Jr., W. F., Vokrouhlick´y, D., Rubincam, D. P., Nesvorn´y, D., 2006. The Yarkovsky and Yorp Effects: Implications for Asteroid Dynamics.
Annual Review of Earth and Planetary Sciences 34, 157–191.
16
Breiter, S., Michalska, H., 2008. YORP torque as the function of shape harmonics. Mon. Not. R. Astron. Soc. 388, 927–944.
Chapman, C. R., Merline, W. J., Thomas, P., 1999. Cratering on Mathilde. Icarus 140, 28–33.
Chapman, C. R., Merline, W. J., Thomas, P. C., Joseph, J., Cheng, A. F., Izenberg, N., 2002. Impact History of Eros: Craters and Boulders. Icarus
155, 104–118.
Chapman, C. R., Ryan, E. V., Merline, W. J., Neukum, G., Wagner, R., Thomas, P. C., Veverka, J., Sullivan, R. J., 1996a. Cratering on Ida. Icarus
120, 77–86.
Chapman, C. R., Veverka, J., Belton, M. J. S., Neukum, G., Morrison, D., 1996b. Cratering on Gaspra. Icarus 120, 231–245.
Chesley, S. R., Ostro, S. J., Vokrouhlicky, D., Capek, D., Giorgini, J. D., Nolan, M. C., Margot, J.-L., Hine, A. A., Benner, L. A. M., Chamberlin,
A. B., 2003. Direct detection of the yarkovsky effect by radar ranging to asteroid 6489 golevka. Science 302 (5651), 1739–1742.
Durech, J., Kaasalainen, M., Marciniak, A., Allen, W. H., Behrend, R., Bembrick, C., Bennett, T., Bernasconi, L., Berthier, J., Bolt, G., Boroumand,
S., Crespo da Silva, L., Crippa, R., Crow, M., Durkee, R., Dymock, R., Fagas, M., Fauerbach, M., Fauvaud, S., Frey, M., Gonc¸alves, R., Hirsch,
R., Jardine, D., Kami´nski, K., Koff, R., Kwiatkowski, T., L´opez, A., Manzini, F., Michałowski, T., Pacheco, R., Pan, M., Pilcher, F., Poncy, R.,
Pray, D., Pych, W., Roy, R., Santacana, G., Slivan, S., Sposetti, S., Stephens, R., Warner, B., Wolf, M., 2007. Physical models of ten asteroids
from an observers' collaboration network. Astron. Astrophys. 465, 331–337.
Durech, J., Vokrouhlick´y, D., Kaasalainen, M., Higgins, D., Krugly, Y. N., Gaftonyuk, N. M., Shevchenko, V. G., Chiorny, V. G., Hamanowa,
H., Hamanowa, H., Reddy, V., Dyvig, R. R., 2008a. Detection of the YORP effect in asteroid (1620) Geographos. Astron. Astrophys. 489,
L25–L28.
Durech, J., Vokrouhlick´y, D., Kaasalainen, M., Weissman, P., Lowry, S. C., Beshore, E., Higgins, D., Krugly, Y. N., Shevchenko, V. G., Gaftonyuk,
N. M., Choi, Y.-J., Kowalski, R. A., Larson, S., Warner, B. D., Marshalkina, A. L., Ibrahimov, M. A., Molotov, I. E., Michałowski, T., Kitazato,
K., 2008b. New photometric observations of asteroids (1862) Apollo and (25143) Itokawa - an analysis of YORP effect. Astron. Astrophys.
488, 345–350.
Gaskell, R., Saito, J., Ishiguro, M., Kubota, T., Hashimoto, T., Hirata, N., Abe, S., Barnouin-Jha, O. S., Scheeres, D., 2006. Global Topography of
Asteroid 25143 Itokawa. In: Mackwell, S., Stansbery, E. (Eds.), 37th Annual Lunar and Planetary Science Conference. Vol. 37 of Lunar and
Planetary Institute Conference Abstracts. pp. 1876–1877.
Greenberg, R., Nolan, M. C., Bottke, Jr., W. F., Kolvoord, R. A., Veverka, J., 1994. Collisional history of Gaspra. Icarus 107, 84–97.
Hapke, B., 1984. Bidirectional reflectance spectroscopy. III - Correction for macroscopic roughness. Icarus 59, 41–59.
Hapke, B., 2002. Bidirectional Reflectance Spectroscopy. 5. The Coherent Backscatter Opposition Effect and Anisotropic Scattering. Icarus 157,
523–534.
Harris, A. W., Fahnestock, E. G., Pravec, P., Feb. 2009. On the shapes and spins of "rubble pile" asteroids. Icarus 199, 310–318.
Helfenstein, P., Veverka, J., 1989. Physical characterization of asteroid surfaces from photometric analysis. In: Binzel, R. P., Gehrels, T., Matthews,
M. S. (Eds.), Asteroids II. pp. 557–593.
Holsapple, K. A., 2007. Spin limits of Solar System bodies: From the small fast-rotators to 2003 EL61. Icarus 187, 500–509.
Housen, K. R., Holsapple, K. A., 2003. Impact cratering on porous asteroids. Icarus 163 (1), 102–119.
Kaasalainen, M., 2004. Physical models of large number of asteroids from calibrated photometry sparse in time. Astron. Astrophys. 422, L39–L42.
Kaasalainen, M., Kwiatkowski, T., Abe, M., Piironen, J., Nakamura, T., Ohba, Y., Dermawan, B., Farnham, T., Colas, F., Lowry, S., Weissman, P.,
Whiteley, R. J., Tholen, D. J., Larson, S. M., Yoshikawa, M., Toth, I., Velichko, F. P., 2003. CCD photometry and model of MUSES-C target
(25143) 1998 SF36. Astron. Astrophys. 405, L29–L32.
Kaasalainen, M., Lamberg, L., Lumme, K., Bowell, E., 1992. Interpretation of lightcurves of atmosphereless bodies. I - General theory and new
inversion schemes. Astron. Astrophys. 259, 318–332.
Kaasalainen, M., Torppa, J., 2001. Optimization Methods for Asteroid Lightcurve Inversion. I. Shape Determination. Icarus 153, 24–36.
Kaasalainen, M., Durech, J., Warner, B. D., Krugly, Y. N., Gaftonyuk, N. M., 2007. Acceleration of the rotation of asteroid 1862 Apollo by
radiation torques. Nature 446, 420–422.
Lagerros, J. S. V., 1996. Thermal physics of asteroids. I. Effects of shape, heat conduction and beaming. Astron. Astrophys. 310, 1011–1020.
Lamy, P. L., Kaasalainen, M., Lowry, S., Weissman, P., Barucci, M. A., Carvano, J., Choi, Y.-J., Colas, F., Faury, G., Fornasier, S., Groussin, O.,
Hicks, M. D., Jorda, L., Kryszczynska, A., Larson, S., Toth, I., Warner, B., 2008. Asteroid 2867 steins. Astronomy and Astrophysics 487 (3),
1179–1185.
Lowry, S. C., Fitzsimmons, A., Pravec, P., Vokrouhlick´y, D., Boehnhardt, H., Taylor, P. A., Margot, J.-L., Gal´ad, A., Irwin, M., Irwin, J., Kusnir´ak,
P., 2007. Direct Detection of the Asteroidal YORP Effect. Science 316, 272–.
Micheli, M., Paolicchi, P., 2008. YORP effect on real objects. I. Statistical properties. Astron. Astrophys. 490, 387–391.
Muinonen, K., Lagerros, J. S. V., 1998. Inversion of shape statistics for small solar system bodies. Astron. Astrophys. 333, 753–761.
Mysen, E., 2008a. An analytical model for YORP and Yarkovsky effects with a physical thermal lag. Astron. Astrophys. 484, 563–573.
Mysen, E., 2008b. Dynamical effects of thermal emission on asteroids. Mon. Not. R. Astron. Soc. 383, L50–L53.
Nakamura, A. M., 2002. Cratering of asteroids and small bodies. Advances in Space Research 29 (8), 1221–1230.
Nesvorn´y, D., Vokrouhlick´y, D., 2007. Analytic theory of the yorp effect for near-spherical objects. The Astronomical Journal 134 (5), 1750–1768.
Nesvorn´y, D., Vokrouhlick´y, D., 2008. Analytic theory for the yarkovsky-o'keefe-radzievski-paddack effect on obliquity. The Astronomical Journal
136 (1), 291–299.
Nesvorn´y, D., Vokrouhlick´y, D., 2008. Vanishing torque from radiation pressure. Astron. Astrophys. 480, 1–3.
O'Brien, D. P., Greenberg, R., Richardson, J. E., 2006. Craters on asteroids: Reconciling diverse impact records with a common impacting
population. Icarus 183 (1), 79–92.
Ostro, S. J., Benner, L. A. M., Nolan, M. C., Magri, C., Giorgini, J. D., Scheeres, D. J., Broschart, S. B., Kaasalainen, M., Vokrouhlick´y, D.,
Chesley, S. R., Margot, J.-L., Jurgens, R. F., Rose, R., Yeomans, D. K., Suzuku, S., de Jong, E. M., 2004. Radar observations of asteroid 25143
Itokawa (1998 SF36). Meteoritics and Planetary Science 39, 407–424.
Rubincam, D. P., 2000. Radiative Spin-up and Spin-down of Small Asteroids. Icarus 148, 2–11.
Saito, J., Miyamoto, H., Nakamura, R., Ishiguro, M., Michikami, T., Nakamura, A. M., Demura, H., Sasaki, S., Hirata, N., Honda, C., Yamamoto,
A., Yokota, Y., Fuse, T., Yoshida, F., Tholen, D. J., Gaskell, R. W., Hashimoto, T., Kubota, T., Higuchi, Y., Nakamura, T., Smith, P., Hiraoka,
17
K., Honda, T., Kobayashi, S., Furuya, M., Matsumoto, N., Nemoto, E., Yukishita, A., Kitazato, K., Dermawan, B., Sogame, A., Terazono, J.,
Shinohara, C., Akiyama, H., 2006. Detailed Images of Asteroid 25143 Itokawa from Hayabusa. Science 312, 1341–1344.
Scheeres, D. J., 2007a. Rotational fission of contact binary asteroids. Icarus 189, 370–385.
Scheeres, D. J., 2007b. The dynamical evolution of uniformly rotating asteroids subject to YORP. Icarus 188, 430–450.
Scheeres, D. J., Abe, M., Yoshikawa, M., Nakamura, R., Gaskell, R. W., Abell, P. A., 2007. The effect of YORP on Itokawa. Icarus 188, 425–429.
Scheeres, D. J., Gaskell, R. W., 2008. Effect of density inhomogeneity on yorp: The case of itokawa. Icarus 198 (1), 125–129.
Slivan, S. M., Sep. 2002. Spin vector alignment of Koronis family asteroids. Nature 419, 49–51.
Taylor, P. A., Margot, J.-L., Vokrouhlick´y, D., Scheeres, D. J., Pravec, P., Lowry, S. C., Fitzsimmons, A., Nolan, M. C., Ostro, S. J., Benner,
L. A. M., Giorgini, J. D., Magri, C., 2007. Spin Rate of Asteroid (54509) 2000 PH5 Increasing Due to the YORP Effect. Science 316, 274–277.
Thomas, P. C., Joseph, J., Carcich, B., Veverka, J., Clark, B. E., Bell, J. F., Byrd, A. W., Chomko, R., Robinson, M., Murchie, S., Prockter, L.,
Cheng, A., Izenberg, N., Malin, M., Chapman, C., McFadden, L. A., Kirk, R., Gaffey, M., Lucey, P. G., 2002. Eros: Shape, topography, and
slope processes. Icarus 155 (1), 18–37.
Thomas, P. C., Veverka, J., Bell, J. F., Clark, B. E., Carcich, B., Joseph, J., Robinson, M., McFadden, L. A., Malin, M. C., Chapman, C. R., Merline,
W., Murchie, S., 1999. Mathilde: Size, shape, and geology. Icarus 140 (1), 17–27.
Veverka, J., Thomas, P. C., Robinson, M., Murchie, S., Chapman, C., Bell, M., Harch, A., Merline, W. J., Bell, J. F., I., Bussey, B., Carcich, B.,
Cheng, A., Clark, B., Domingue, D., Dunham, D., Farquhar, R., Gaffey, M. J., Hawkins, E., Izenberg, N., Joseph, J., Kirk, R., Li, H., Lucey, P.,
Malin, M., McFadden, L., Miller, J. K., Owen, W. M., J., Peterson, C., Prockter, L., Warren, J., Wellnitz, D., Williams, B. G., Yeomans, D. K.,
2001. Imaging of Small-Scale Features on 433 Eros from NEAR: Evidence for a Complex Regolith. Science 292 (5516), 484–488.
Vokrouhlick´y, D., Capek, D., 2002. YORP-Induced Long-Term Evolution of the Spin State of Small Asteroids and Meteoroids: Rubincam's
Approximation. Icarus 159, 449–467.
Vokrouhlick´y, D., Capek, D., Kaasalainen, M., Ostro, S. J., 2004. Detectability of YORP rotational slowing of asteroid 25143 Itokawa. Astron. As-
trophys. 414, L21–L24.
Walsh, K. J., Richardson, D. C., Michel, P., 2008. Rotational breakup as the origin of small binary asteroids. Nature 454, 188–191.
18
|
1806.01284 | 1 | 1806 | 2018-06-04T18:00:02 | Exocomet Orbit Fitting: Accelerating Coma Absorption During Transits of $\beta$ Pictoris | [
"astro-ph.EP"
] | Comets are a remarkable feature in our night sky, visible on their passage through the inner Solar system as the Sun's energy sublimates ices and liberates surface material, generating beautiful comae, dust, and ion tails. Comets are also thought to orbit other stars, and are the most promising interpretation of sporadic absorption features (i.e. transits) seen in spectra of stars such as $\beta$ Pictoris and 49 Ceti. These `exocomets' are thought to form and evolve in the same way as in the Solar system, and as in the Solar system we may gain insight into their origins by deriving their orbits. In the case of $\beta$ Pictoris, orbits have been estimated indirectly, using the radial velocity of the absorption features coupled with a physical evaporation model to estimate the stellocentric distance at transit $d_{\rm tr}$. Here, we note that the inferred $d_{\rm tr}$ imply that some absorption signatures should accelerate over several hours, and show that this acceleration is indeed seen in HARPS spectra. This new constraint means that orbital characteristics can be obtained directly, and the pericentre distance and longitude constrained when parabolic orbits are assumed. The results from fitting orbits to 12 accelerating features, and a handful of non-accelerating ones, are in broad agreement with previous estimates based on an evaporation model, thereby providing some validation of the exocomet hypothesis. A prediction of the evaporation model, that coma absorption is deeper for more distant transits, is also seen here. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 000–000 (0000)
Printed 6 June 2018
(MN LATEX style file v2.2)
Exocomet Orbit Fitting: Accelerating Coma Absorption During
Transits of β Pictoris
Grant M. Kennedy⋆1,2
1 Department of Physics, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK
2 Centre for Exoplanets and Habitability, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK
6 June 2018
ABSTRACT
Comets are a remarkable feature in our night sky, visible on their passage through the
inner Solar system as the Sun's energy sublimates ices and liberates surface material,
generating beautiful comae, dust, and ion tails. Comets are also thought to orbit other
stars, and are the most promising interpretation of sporadic absorption features (i.e.
transits) seen in spectra of stars such as β Pictoris and 49 Ceti. These "exocomets" are
thought to form and evolve in the same way as in the Solar system, and as in the Solar
system we may gain insight into their origins by deriving their orbits. In the case of β
Pictoris, orbits have been estimated indirectly, using the radial velocity of the absorption
features coupled with a physical evaporation model to estimate the stellocentric distance
at transit dtr. Here, we note that the inferred dtr imply that some absorption signatures
should accelerate over several hours, and show that this acceleration is indeed seen in
HARPS spectra. This new constraint means that orbital characteristics can be obtained
directly, and the pericentre distance and longitude constrained when parabolic orbits
are assumed. The results from fitting orbits to 12 accelerating features, and a handful
of non-accelerating ones, are in broad agreement with previous estimates based on an
evaporation model, thereby providing some validation of the exocomet hypothesis. A
prediction of the evaporation model, that coma absorption is deeper for more distant
transits, is also seen here.
Key words: comets: general - planets and satellites: detection - planetary systems
- planet-disc interactions - circumstellar matter - stars: individual: β Pictoris
1 INTRODUCTION
Comets are a well-known and important component of our Solar
system, and while a scientific focus for astronomers, some are
visible to the naked eye and capture the imagination of millions.
These icy bodies become visibly more active as they near the
Sun, providing a window into the primordial conditions in the
outer Solar system (e.g. Blum et al. 2017). Comets formed be-
yond the Asteroid belt, and are only brought within a few astro-
nomical units through interaction with the giant planets; in the
case of shorter period comets (e.g. Jupiter Family Comets) these
interactions are recent and ongoing (e.g. Duncan & Levison
1997), while in the case of long period comets, these interac-
tions occurred billions of years ago and resulted in the forma-
tion of the Oort cloud, the source of long period comets (e.g.
Duncan et al. 1987).
While our detailed knowledge of planetary systems around
other stars is relatively limited in comparison, it is clear that
processes analogous to those postulated for the Solar sys-
⋆ Email: [email protected]
c(cid:13) 0000 RAS
tem's formation and evolution do occur. Most obviously, plan-
ets exist in spectacular abundance (e.g. Petigura et al. 2013),
and the cometary source regions known as "debris discs" are
seen around at least 30% of other stars (e.g. Eiroa et al. 2013;
Sibthorpe et al. 2018). Of particular relevance here is the fact
that the first clear evidence of another planetary system – the
image of the edge-on disc orbiting β Pictoris (Smith & Terrile
1984) – prompted a series of papers that may mark the dis-
covery of extrasolar comets (e.g. Kondo & Bruhweiler 1985;
Lagrange et al. 1987; Ferlet et al. 1987; Lagrange-Henri et al.
1988). These "exocomets" manifested as transient red-shifted
ionic absorption lines in high-resolution ultraviolet and optical
spectra; the signature of a large coma transiting the face of the
star. The exocomet interpretation was (and perhaps still is) con-
troversial, so these events have also become known by the more
prosaic term "falling evaporating bodies", or FEBs.
These discoveries were backed up by the development of
a physical model, whereby material is ejected from a cometary
nucleus into a coma dominated by neutral hydrogen. Observed
elements such as calcium and magnesium are ionised when vis-
ible to stellar radiation, and are subsequently blown into a ten-
2
Grant M. Kennedy
uous ion tail by radiation pressure, thus forming a parabolic
front of metallic ions around the nucleus (Beust et al. 1989,
1990). The effect of radiation pressure was found to be criti-
cal to explain the observations; the closer the coma to the star,
the stronger the radiation pressure and the smaller the front of
ions, and therefore the shallower the absorption. While subject
to systematic uncertainties from model assumptions, this prop-
erty means that the distance of the coma to the star for a given
transit could be estimated from the depth of the absorption line.
The key modelling conclusion was that the velocity of the ob-
served absorption corresponds to the radial velocity of the coma
(Beust et al. 1990), meaning that by measuring these velocities,
and estimating the distance to the star at transit, one may derive
constraints on the comet orbit.
A common inference from application of the model to
the data for β Pictoris, and indeed which could be inferred
from the data themselves, is that the tendency of the FEBs
to be redshifted implies that the cometary orbits are not ran-
domly distributed. The relative lack of blue shifted events
(Lagrange-Henri et al. 1992; Crawford et al. 1998) means that
most comets are approaching pericentre, with values for the
longitude of pericentre (measured from the line of sight
to the star), being between -50 and 130◦ (Beust et al. 1990;
Kiefer et al. 2014).
This conclusion led to proposals for three main theories
for the origins of β Pictoris' transiting comets. The first is
that the events represent a single family of comets on common
orbits, the aftermath of the breakup of a single larger object
(Beust & Lissauer 1994). The main issue with this theory is that
we are unlikely to witness the aftermath of such an event, sug-
gesting that mechanisms that work on longer timescales would
provide more satisfactory explanations.
One such theory appeals to the Solar system, noting that
the eccentricities of Asteroids whose pericentres precess at the
same rate as Saturn (i.e. those in the ν6 secular resonance) in-
crease in a way that is correlated with Saturn's pericentre. Thus,
these bodies make their closest approaches to the Sun with a
preferred pericentre direction, and the same explanation was of-
fered for β Pictoris' FEBs (Levison et al. 1994). The criticism
levelled at this theory is that it is very specific, and has not
been shown to generically reproduce the FEB phenomenon (see
Beust & Morbidelli 1996).
theory, and the one that has received the
The final
most attention,
is that
the 4:1 and/or 3:1 resonance with
an outer planet lies within a source belt (Beust & Morbidelli
1996; Quillen & Holman 2000; Beust & Morbidelli 2000;
Th´ebault & Beust 2001; Pichierri et al. 2017). The attraction of
this theory is that objects in these resonances can be excited
to star- grazing orbits with a very specific range of pericen-
tre longitudes. A potential weakness of this theory (which also
applies to secular resonances) is that the timescale to reach
very high eccentricity orbits may be long relative to the time
taken to destroy the comet by evaporation, limiting the ability
of the resonance to produce the wide range of pericentre dis-
tances that is inferred from the models described above. Vari-
ous solutions to this issue exist. One is that some comets are
larger than about 30 km in radius, so can survive long enough
to reach small pericentre distances (Beust & Morbidelli 2000).
Another invokes the existence of additional planets that either
perturb the first planet, or perturb the comets directly, thus caus-
ing larger orbit to orbit changes in the comet's pericentre dis-
tances (Beust & Morbidelli 1996, 2000). One of the ways that
this theory has been tested and developed is by aiming to match
the derived properties of FEBs, for example their distributions
of periastron distances, pericentre angles, and stellocentric dis-
tances at transit (Beust & Morbidelli 2000).
2 MOTIVATION
Thousands of spectra of β Pictoris have been taken in pur-
suit of FEBs, but one aspect that has received little attention
is the prospect for directly deriving the cometary orbits. While
orbital characteristics have been derived based on the evapo-
ration theory (Beust et al. 1990; Kiefer et al. 2014), these are
model-dependent. That is, while the evaporation theory clearly
reproduces the observations very well, it necessarily includes
many assumptions for quantities that may vary from comet to
comet, and it would be far preferable to derive the orbits in a
less model-dependent way. Such an ability would provide fur-
ther tests of the exocomet hypothesis, and more certain tests of
comet origin models. Orbital constraints mean that the evapora-
tion model could be refined, leading to a better physical under-
standing of the comets and their comae.
To derive an orbit from the absorption lines, enough infor-
mation to constrain the orbital elements is needed (here we use
pericenter distance q, eccentricity e, inclination i, line of nodes
Ω, longitude of pericentre , and true anomaly f ). All six are
not necessary however, because we may assume an edge-on or-
bit, thus eliminating i and Ω. We may further eliminate f by
defining to be the angle from the line of sight to the star at
transit centre (i.e. f = −). An assumption, that the orbits
are parabolic (e = 1) may be made based on the expectation
that comets cannot survive more than a few periastron passages
near β Pictoris before evaporating. That is, the comets are not
spiraling in on near-circular orbits, but are appearing at just a
few stellar radii via relatively large orbit-to-orbit changes in q
(Beust & Morbidelli 1996). At this point, two further indepen-
dent pieces of information are needed to constrain the orbit. In
previous work this has been the observed radial velocity of the
absorption line, and the distance to the star at transit estimated
from the evaporation model.
Given that FEBs are estimated to lie within a few tens of
stellar radii when they are observed, one may ask whether accel-
eration could be detected in the course of a night's observations.
For an object at 10R⋆ from β Pictoris (using R⋆ = 1.5R⊙ and
M⋆ = 1.7M⊙), the acceleration due to gravity is 2 m s−2, or
7 km s−1 h−1. The magnitude of this acceleration is within easy
reach of high-resolution spectrographs such as the High Ac-
curacy Radial velocity Planet Searcher (HARPS), which has a
spectral resolution equivalent to a few km s−1. Therefore, given
that absorption features inferred to be caused by absorption of
material at less than ten stellar radii are regularly seen, the ex-
pected accelerations should be present and readily detectable in
sequences of spectra spanning more than about one hour. Non-
detection of acceleration where sufficient temporal coverage ex-
ists would be a serious issue for the exocomet hypothesis.
Aside from the expectation of acceleration given the likely
orbits, hints that it should be detectable can be seen in previ-
ous works; Ferlet et al. (1987) show a series of spectra taken
over several hours, in which a redshifted line clearly acceler-
ates by about 5 km s−1, and temporal variation was also seen
by Lagrange et al. (1996) and Petterson & Tobin (1999) With
the detection of such an acceleration, the distance derived from
c(cid:13) 0000 RAS, MNRAS 000, 000–000
x
u
l
f
d
e
s
i
l
a
m
r
o
n
x
u
l
f
d
e
s
i
l
a
m
r
o
n
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.6
0.4
0.2
0.0
−40
−20
Ca II K
Ca II H
60
80
100
20
0
radial velocit / km s−1
40
Figure 1. Examples of five HARPS spectra, showing the calcium K
and H lines centred at the radial velocity of β Pictoris. The dashed line
shows the stellar reference spectrum. The absorption at zero km s−1
is thought to originate in a stable circumstellar disc, while the varying
components are attributed to exocomet comae. Features at both low and
high velocity are seen, with those at low velocity tending to be deeper.
the evaporation model would no longer be needed to constrain
a parabolic orbit (or could be used to refine the evaporation
model).
3 DETECTION OF EXOCOMET ACCELERATION
As a nearby (19pc) naked eye A6V star (V = 3.8), β Pictoris is
observable at high signal to noise ratios with a variety of instru-
ments. One particular focus has been high resolution (∆λ/λ ∼
105) spectra, which in addition to FEBs, reveals a relatively
stable circumstellar calcium absorption line (e.g. Hobbs et al.
1985) and stellar pulsations (Koen et al. 2003), and yields con-
straints on the possible orbits of unseen planets (Galland et al.
2006) and on the mass of β Pictoris b (Lagrange et al. 2012).
Specifically, the latter study obtained about 1000 HARPS
spectra between 2004 and 2011. In the time since, observations
have continued, and in late 2017 all 2642 public HARPS spec-
tra of β Pictoris were downloaded from the ESO archive for use
in this study. These spectra have been processed by the HARPS
Data Reduction Software, and have been shown by Kiefer et al.
(2014) to be of sufficient quality for work related to FEBs; the
spectra have a stable wavelength calibration that far exceeds our
needs here, and the relative flux calibration across the spectral
lines of interest and over the years since 2004 is high enough
that variations at the ∼1% level are easily detected (see for ex-
ample their Extended Data Fig. 2).
Here, we focus on the calcium K & H lines at 3933.66
and 3968.47 A. The only post-processing applied to the spectra
here is to extract the region within ±500 km s−1 of the calcium
lines, and apply a multiplicative normalisation in the wings of
the line so that all spectra are on the same (arbitrary) flux scale.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Exocomet Orbit Fitting
3
At this stage a "stellar" reference spectrum near each line is
computed; the method follows very closely that described by
Kiefer et al. (2014), which used the highest flux values minus
an estimated noise level across all spectra to derive a spectrum
free of absorption. An additional step added here is interpola-
tion across the circumstellar line, which is always present but
varies in width/velocity slightly, so that reference divided (or
subtracted) spectra do not show spurious variations that might
be interpreted as FEB activity near zero velocity.
An example of five "randomly" chosen spectra is shown
in Figure 1, where the stellar systemic velocity has been set to
21 km s−1.1 The obvious feature of all spectra is that the cal-
cium lines are broadened by a few hundred km s−1 by the fast
stellar rotation, and in addition every spectrum shows the stable
circumstellar line and several other narrow (few to a few tens of
km s−1) and/or broad (few tens to hundreds of km s−1) absorp-
tion features; the FEBs. The stable circumstellar line is present
in all spectra, and is seen to move and/or change in width at
the ∼km s−1 level. Spectra generally appear different night-to-
night, because the transit time for comets is of order hours. Ab-
sorption features are generally present in both Ca lines, though
are deeper and more apparent in K than H unless the absorption
is optically thick (i.e. the optical depth can be estimated from the
line ratios, e.g. Lagrange-Henri et al. 1992; Petterson & Tobin
1999).
To detect accelerations we start with the graphical tech-
nique of plotting temporal sequences of spectra as images, an
example of which is shown in the upper panel of Figure 2. The
spectral dimension is simply that output by the HARPS reduc-
tion pipeline, but spectra are binned temporally; each vertical
pixel is 1/250th of a day (about 6 minutes), so rows where mul-
tiple spectra fall are averaged. On nights with continuous moni-
toring, bins typically contain five spectra. A weak temporal vari-
ation is present on most nights, with a period of approximately
0.6 hours; this has been removed by subtracting a smoothed im-
age with the absorption lines removed (see section 4). A final
feature present is pulsation, whose primary signal is diagonal
stripes with a gradient of approximately 120 km s−1 h−1; it is
strongest outside the deepest part of the lines (&200 km s−1
at about the 1% level) so not visible in Figure 2, and not cor-
rected for. The pulsation signal is much weaker than the accel-
erating features considered here, and has a greater velocity gra-
dient than all fitted features, so is unlikely to affect the results.
An alternative way to visualise the results is simply to plot
the spectra, as is done in the lower panel of Figure 2. This panel
shows the same spectra as in the upper panel, but binned tempo-
rally by a factor of two, and spectrally by a factor of four. These
spectra can highlight temporal variations of absorption lines by
creating the illusion of a surface; static absorption lines result
in a series of features that are directly above each other (i.e. a
vertical 'valley'), while those that accelerate show features that
move in velocity with each successive spectrum (i.e. a diagonal
valley).
Figure 2 shows three absorption features, two at 0 and 25
km s−1 that do not accelerate, and one at 60-80 km s−1 that
does. The accelerating feature is seen as either a stripe in the
upper panel, or a rightward-moving dip in the lower panel.
1 i.e. the spectra shown were not chosen completely randomly, but only
a few attempts need to be made to obtain a similarly varied set of spectra
where both deep and blueshifted lines are present.
4
Grant M. Kennedy
2
0
.
J
2
4
5
4
5
=
D
M
m
o
r
f
s
r
u
o
h
1.75
1.50
1.25
1.00
0.75
0.50
0.25
0.00
1.2
ln(1-flux/reference)
t
e
f
f
o
+
e
c
n
e
r
e
f
e
r
/
x
u
l
f
1.0
0.8
0.6
0.4
0.2
0.0
−100
−50
0
50
100
radial velocity / km −1
150
200
250
Figure 2. Example of an accelerating Ca K line absorption feature, seen at 60-80 km s−1. The upper panel shows the level of absorption on a log
stretch, and the lower panel shows 1d spectra that have been divided by the reference and offset vertically for clarity. The 1d spectra have been binned
by a factor of two in time, and four in velocity, relative to the image in the upper panel. Two strong lines at vr ≈ 0 and 25 km s−1, plus a weaker one
moving from 60 to 80 km s−1, are visible, and the latter is the accelerating absorption feature. In the image this moving feature is seen as a faint stripe
moving up and to the right over time. In the series of spectra the feature appears as a series of small absorptions that move up and to the right over time.
Both panels illustrate clearly that this absorption feature is moving over the observation sequence.
A set of accelerating features were identified in spectra by
eye using images similar to Figure 2, from 35 nights' data with
more than about an hour of continuous observation, and these
are shown in Figures 3 and 4. The dates on which these se-
quences started are shown in each panel, and given in Table 1.
In most cases the accelerating feature is apparent, though the
relatively low contrast relative to the deeper circumstellar line
makes their visualisation difficult in some cases. Also, in a few
cases the acceleration is not large enough to be clearly visible
(e.g. the 40 km s−1 feature in second panel from top is accelerat-
ing slightly, but significantly). While not all features are clearly
visible in the figures, the identified features are verified to i)
have average absorption significantly greater than zero (i.e. ex-
ist), and ii) be accelerating significantly, by the fitting method
outlined in section 4.
All accelerating features found are red shifted, and are ac-
celerating towards the star. No automated detection methods
were attempted, though we found that some features could also
be discovered by fitting Gaussians to absorption features in in-
dividual spectra, and plotting the temporal evolution of their
radial velocities (see Petterson & Tobin 1999, for such plots).
In addition, distinct unblended absorption features that did not
show apparent acceleration were also selected for further anal-
ysis, with the expectation that useful orbital constraints could
still arise (e.g. a minimum distance at transit). For these, only
red shifted events were selected because the aim is to illustrate
orbital fitting of accelerating features, and to compare the orbital
constraints for both accelerating and non-accelerating features.
There is a degree of subjectivity in the subset of spectra used
below, and more accelerating features might be discovered with
a more systematic approach. We are not however attempting to
make any statistical inference from the subset of features anal-
ysed, so any biases introduced by our approach do not influence
the results that follow.
4 ORBIT FITTING
Given an accelerating absorption feature it is relatively simple
to construct and fit a model of an orbit. The primary assump-
tion here is that the orbit crosses the centre of the stellar disc,
leaving three elements to be constrained (eccentricity e, longi-
tude of pericenter , and pericentre distance q). Thus, an or-
bital fit will yield constraints, but with a strong degeneracy be-
tween these within the allowed parameter space. However, as
discussed above the reasonable assumption of a parabolic or-
bit (e = 1), means that an accelerating feature contains enough
information to constrain the orbit fully. Here we illustrate this
degeneracy by leaving e to vary, but also restrict the results to
parabolic orbits.
Radial velocity vr is the variable being fitted here, with
respect to Earth once β Pictoris' systemic velocity has been
subtracted (i.e. vr is only the same as the stellocentric radial
velocity at transit center). This velocity can be derived using
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Exocomet Orbit Fitting
5
s
r
u
o
h
/
e
m
i
t
1.5
1.0
0.5
0.0
t
e
s
f
f
o
+
e
c
n
e
r
e
f
e
r
/
x
u
l
f
1
0
54542
54829
54913
55170
55567
55597
56685
56694
56982
56988
57344
−50
0
50
150
radial velocity / km s−1
100
200
250
−50
0
50
150
radial velocity / km s−1
100
54542
54829
54913
55170
55567
55597
56685
56694
56982
56988
57344
200
250
Figure 3. Accelerating features identified for modelling, indicated by
transparent dashed lines. The modified Julian date at the start of each ob-
serving sequence is given. Each panel is the same as the upper panel in
Figure 2 and has the same scale, where lighter colours indicate stronger
absorption.
Figure 4. Spectra of accelerating features identified for modelling. The
modified Julian date at the start of each observing sequence is given.
Each panel is the same as the lower panel in Figure 2. As discussed in
the text, weak accelerating features are not always easily discernible.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
6
Grant M. Kennedy
the motion of an orbit in a frame x, y, where the pericentre lies
along the positive x coordinate, and β Pictoris is at the origin.
The system is viewed from the negative Y direction, where the
X, Y coordinates are rotated clockwise by from x, y (e.g.
Murray & Dermott 2000). That is, the orbit in the x, y frame is
rotated such that the pericentre direction is at an angle from
−Y (the reader may wish to refer ahead to Fig. 6.) In the frame
of the orbit
x = −µ/h sin f
y = µ/h(1 + cos f )
(1)
where µ = GM⋆, f is the true anomaly, h2 = µp, and p = 2q
for parabolic orbits and p = a(1 − e2) for elliptical orbits. The
rotation yields
vr = Y = − x cos + y sin
= µ/h (sin f cos + sin [1 + cos f ])
(2)
We fit absorption features directly (i.e. instead of fitting ac-
celerations and then deriving orbital constraints), so deriving an
orbit from the images requires five or six free parameters, the
orbital elements q, , and e (for elliptical orbits), the time t0 of
mid-transit relative to the start of the observation sequence, and
the (constant) depth δ and velocity width σ of the absorption
line, which is assumed to be Gaussian. The radial velocity at a
given time t is found by first finding the true anomaly f (t) (us-
ing Barker's equation in the case of parabolic orbits), and then
using equation (2). Thus, a model of an absorption line analo-
gous to the accelerating feature in Figure 2 is created by com-
puting vr(t) at the centre of each temporal bin, and the image
comprises a vertical sequence of (negative) Gaussians of depth
δ (where positive values mean absorption) and width σ centred
at vr(t). At each time, an absorption line is only assumed to
be present if the centre of the body is in front of the star (i.e.
X < R⋆). Additional parameters that can be derived from the
models2 are the radial velocity at mid-transit vr,mid−transit, the
distance to the star at mid-transit dtr (= −Y ), and the accelera-
tion vr = µ/d2
tr.
As can be seen from Figures 3 and 4, most spectra con-
tain fixed absorption lines that are much deeper than the ac-
celerating ones. Fitting the model and obtaining reasonable χ2
values from residual images that are also easily interpreted vi-
sually therefore requires that these are removed. First, we find
the median spectrum over the nightly sequence. The noise level
in each temporal bin is then computed as the standard deviation
of the residuals when this median is subtracted from the initial
image. Next, a second order polynomial is fitted to the median
in the region near the line of interest (typically 20 km s−1 on ei-
ther side, though in cases with other nearby lines (e.g. for lines
with low vr) this region is necessarily smaller). The outer edges
of this region have twice the weight relative to the centre in
the fit, and this polynomial is taken as an estimate of the local
time-independent background near the accelerating absorption
feature. This polynomial fit to the median was found to be much
better than simply using the median, as the absorption line of in-
terest was commonly partially subtracted with the latter method.
During the fitting we subtract this polynomial and the proposed
model from the image to obtain a residual image.
2 We do not distinguish between dtr and dtr,mid−transit because the
distance to the star changes little during a transit. We use vr instead of
vr,mid−transit for the same reason.
Fitting orbits to the spectra is done in two steps, where the
goodness of fit metric is the sum of squares of the residuals di-
vided by the noise. First, a model is fitted to the original image
with the polynomial subtracted to obtain a preliminary fit. A
preliminary residual image is created by subtracting the poly-
nomial and this model, and the residuals are smoothed with a
small temporal extent of 8.5 minutes, but a wide spectral extent
of 100 km s−1. These smoothed residuals (which vary at the
percent level relative to the absorption features) are then sub-
tracted from the original image before fitting.
As noted above, very strong degeneracies between fit-
ted parameters exist when the eccentricity is not fixed, so
the second step is to use a Markov Chain Monte Carlo
(MCMC) method to explore the parameter space. Specifi-
cally, we use the ensemble sampler emcee (Goodman & Weare
2010; Foreman-Mackey et al. 2013) with 256 parallel chains
("walkers"). The chains are run for 100,000 steps to ensure the
parameter space is fully explored, as in some cases the autocor-
relation lengths were of order 10,000 steps. The only restrictions
on the fitted parameters are that orbits are bound or parabolic
(i.e. e 6 1), do not hit the star at pericentre (i.e. q > R⋆), and
that the depth δ and width σ are positive.
After these runs, the walkers from the final step were used
to generate distributions of orbital elements and other derived
quantities of interest. This fitting was run twice; once for ellip-
tical orbits, and then again for parabolic orbits, thus obtaining
two sets of orbital elements from each modelled absorption line.
The parabolic orbits are of course a subset of the elliptical or-
bits, but fitting them is more satisfactory than simply applying
an arbitrary cut in eccentricity to the elliptical orbit results (at
e = 0.95, say).
An example of a model fit is shown in Figure 5, for the
same feature in Figure 2 (and the top row of Figure 3). While
the accelerating feature is not subtracted perfectly, the model ad-
equately reproduces the accelerating absorption line, and while
the depth of the absorption can appear to vary (Beust & Lissauer
1994), any variation is not at a sufficiently strong level that
the addition of another model parameter is warranted for the
purposes of orbital fitting. This procedure was applied to all
eleven sets of spectra shown in Figure 3, plus others where non-
accelerating lines were well-separated from others, thus obtain-
ing sets of elliptical and parabolic orbital elements from a set
of 27 absorption features. Of these, twelve are accelerating fea-
tures (there are two in the second panel from the top in Figure
3).
5 RESULTS
The results of the orbital fits may be visualised in various ways;
Figure 6 shows a top-down (X, Y ) view of the final set of orbits
for the example above. From this figure the degeneracies are
clear. A wide range of orbits is possible, and those with smaller
pericentre distances reach pericentre at a greater longitude from
the line of sight to the star. The degeneracy with eccentricity is
more subtle and harder to discern here, as the most eccentric
orbits are in fact those with the greatest pericentre distances and
smallest pericentre longitudes.
These orbits may also be viewed in parameter space plots,
each of which shows clusters of 256 points from the final step in
the MCMC chains. These clusters therefore show the distribu-
tion of possible parameters for a given absorption feature. The
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Exocomet Orbit Fitting
7
ln(1-flux/reference)
Figure 6. Distribution of possible elliptical orbits for the example in
Figure 2. Each line shows one of the 256 walkers at the final step in the
MCMC chains, and each has a corresponding circle at pericenter. The
axes here are X, Y , so pericentre is at an angle anticlockwise from
−Y . The thick black circle in the centre is the star, and the system is
viewed from below (i.e. −Y ).
1
−
h
1
−
s
m
k
/
r
v
102
101
100
10−1
10−2
20
40
80
60
100
vr, mid − transit / km s−1
120
140
160
free-fall
Figure 7. Accelerations for fitted absorption features, plotted against
the distance to the star at mid-transit. Each cluster of dots shows the
final step of the 256 walkers in the MCMC chain for the fit to a given
absorption feature. The dot colours are used consistently across Figs.
7-10. The dashed line shows the acceleration expected for bodies in
free-fall from infinity. Those without detected acceleration extend down
to the free-fall line, and their symbols are transparent.
2
0
.
J
2
4
5
4
5
=
D
M
m
o
r
f
s
r
u
o
h
1.75
1.50
1.25
1.00
0.75
0.50
0.25
0.00
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
t
e
f
f
o
+
e
c
n
e
r
e
f
e
r
/
x
u
l
f
−20
0
20
40
80
radial velocity / km −1
60
data
data-model
100
120
Figure 5. Fitting results, showing the same data as in Figure 2, but with
the best-fit model subtracted. The upper panel shows the level of ab-
sorption after model subtraction, where brighter colours indicate deeper
absorption. The lower panel shows the binned spectra, with the black
lines showing the original data, and the grey lines the model subtracted
data.
first example is Figure 7, which illustrates the distinction be-
tween absorption features that do and do not show measurable
accelerations. While features with significant acceleration yield
a direct measurement of the distance to the star, features with
undetectable accelerations are consistent with orbits that pass
in front of the star at arbitrarily large distances. However, the
restriction that fitted orbits are bound or parabolic means that
for a given velocity at mid-transit the smallest possible accel-
eration coresponds to a object in free-fall from infinity (which
is vr = v4
r /(4µ)). Thus, non- accelerating features are easily
identified in Figure 7 by their lack of clustering at a given vr,
and the extension of their distributions down to the expected
limit based on the fitting method.
In the left panel of Figure 8, all fitted orbits are shown,
and again the distinction is made between orbits with and with-
out detected acceleration. It is clear that degeneracies apply in
all cases. The same trends visible in Figure 6 can be identified,
but it is now apparent that each fit reaches a minimum eccen-
tricity that is consistent with the data. This can be understood
from equation (2), because at transit f ≈ −, and therefore
vr,transit ∝ sin and the maximum radial velocity at transit
for a given orbit occurs when is near 90◦. Thus, the lowest
eccentricity orbit consistent with the data should also be near
90◦. The eccentricities rise to larger pericentre longitudes, but
stop short of e = 1 because of the restriction that the pericentre
distance be larger than R⋆. In cases with detected acceleration e
and are strongly correlated, but this correlation is not present
c(cid:13) 0000 RAS, MNRAS 000, 000–000
ij
8
Grant M. Kennedy
r
a
t
s
R
/
q
102
101
100
1.0
0.8
e
0.6
0.4
0.2
50
100
ϖ / degrees
150
100
101
102
q / Rstar
r
a
t
s
R
/
q
40
35
30
25
20
15
10
5
0
20
30
40
50
60
ϖ / degrees
70
80
90
Figure 8. Distributions of possible orbital elements for all fitted features. Each cluster of dots shows the final step of the 256 walkers in the MCMC
chain for the fit to a given absorption feature. Those without detected acceleration are transparent and dot colours are used consistently across Figs.
7-10. The left panel shows results for elliptical orbits, and the right panel shows results for parabolic orbits. The range in the right panel is restricted
(to the grey box in left panel), as the poorly constrained parabolic orbits look very similar to the elliptical ones. For elliptical orbits the orbital element
degeneracies are strong, but the elements are well constrained for parabolic orbits when acceleration is seen.
for those without and the distributions form a broad cloud of
points centered near ≈ 100◦.
Restricting the orbits to be parabolic yields tighter con-
straints, as shown in the right panel of Figure 8. The plotted
range is smaller here, because non-accelerating features yield
very similar constraints to those for elliptical orbits. For accel-
erating features however, the constraints localise the orbits to
relatively small regions in q − space. A summary of the pa-
rameters derived for parabolic orbits is given in Table 1.
A final way to view the orbit distributions is shown in Fig-
ure 9, primarily for comparison with the models presented in
Beust & Morbidelli (2000) (e.g. see their Figure 4). Here the
radial velocity at the time of mid-transit is used; while Figure 3
shows that in general the velocity of an accelerating feature at
any given time is well constrained, this is not necessarily true at
the time of mid-transit (which may occur during, before, or after
the observation sequence). Thus, as shown in the left panel, for
elliptical orbits the distribution of vr,mid−transit can be rather
wide. When the orbits are restricted to be parabolic, the right
panel shows that the constraints become tighter, though not so
much that this assumption yields significantly more informa-
tion than when eccentricity is a free parameter. That is, while
parabolic orbits must be assumed to obtain tight constraints on
orbital elements, this is less true in the parameter space of Fig-
ure 9.
Non-accelerating features result in poor constraints in Fig-
ure 9, most of which extend beyond the right side of the plot.
The distances at transit are only constrained to be larger than
some value that depends on the strength of the signal (i.e. the
upper limit on the level of acceleration that was not detected).
However, the fact that the absorption features are seen at all
means that dtr is no more than a few tens of stellar radii for
evaporation to be occuring (Beust et al. 1996, derive ∼35R⋆),
so the true dtr of these orbits is likely below this value, and
consistent with the orbits with stronger constraints.
In Figure 9 the general trend is that orbits that transit at
smaller stellocentric distances have a greater radial velocity.
The grey boxes show the estimated regions in which previously
identified populations of FEBs lie, where the evaporation model
was used to estimate the distance at transit, and the distinction
between "high" and "low velocity features" (HVFs and LVFs) is
largely historic.3 The first conclusion is therefore that the orbits
derived here in general have similar properties to those derived
using the evaporation model.
While our orbits tend to lie within the HVF box, they are
on average above the LVFs. However, the lack of orbits falling
within the LVF box is probably the result of a bias, as features
that are the result of transits at greater distances more com-
monly occur at low velocity, and these are more common and
more likely to be blended with other features. Thus, the main
conclusions from placing the orbits derived here on this plot
is that there is not strong evidence for a lack of orbits within
the HVF and LVF boxes, but that orbits certainly exist outside
them, in particular at greater velocities for a given distance at
transit when dtr > 10R⋆. The existence of such orbits is indeed
predicted by the resonance model (Beust & Morbidelli 2000),
and the orbits derived from this method will provide useful new
constraints for this model.
The last view of the fitting results includes the absorption
depth derived in each case, shown against the distance at tran-
sit in Figure 10. The values for δ are in units of normalised
3 Here the 'very low velocity features' (VLVFs) are not considered, as
none in this class were found to be accelerating.
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Exocomet Orbit Fitting
9
HVFs
160
140
120
100
80
60
40
20
0
1
−
s
m
k
/
t
i
s
n
a
r
t
−
d
m
i
,
r
v
LVFs
HVFs
160
140
120
100
80
60
40
20
0
LVFs
0
10
20
dtr / Rstar
30
40
0
10
20
dtr / Rstar
30
40
Figure 9. Distributions of velocity versus stellocentric distance at mid-transit, for elliptical (left panel) and parabolic (right panel) orbits. Each cluster of
dots shows the final step of the 256 walkers in the MCMC chain for the fit to a given absorption feature. Those without detected acceleration are trans-
parent and dot colours are used consistently across Figs. 7-10. The grey boxes show 'high' and 'low velocity features', following Beust & Morbidelli
(2000). The constraints in this space are only moderately improved when parabolic orbits are assumed.
0.30
0.25
0.20
0.15
0.10
0.05
Δ
0.00
100
101
102
dtr / Rstar
103
Figure 10. Correlation between distance at transit and absorption frac-
tion for elliptical orbits. Each cluster of dots shows the final step of
one of the 256 walkers in the MCMC chain for the fit to a given absorp-
tion feature. Those without detected acceleration are transparent and dot
colours are used consistently across Figs. 7-10. This link is predicted by
the evaporation model because exocomet comae are more strongly ab-
lated closer to the star, and therefore cover a smaller fraction of the star.
The constraints are not improved with the assumption of parabolic or-
bits (so that plot is omitted here).
c(cid:13) 0000 RAS, MNRAS 000, 000–000
flux (i.e. as in Figure 1), but can be converted to fractional peak
absorption ∆ using the reference spectrum with ∆ = δ/Fref ,
where Fref is the reference (stellar) flux at rv,mid−transit. The
relevance of this plot is that the evaporation model described
above predicts that comet absorption is deeper for more distant
transits, because the effect of stellar radiation pressure on the
calcium ions is less. Thus the comet comae are larger and cover
a greater fraction of the projected stellar surface. This prediction
is borne out by Figure 10, which shows a clear tendency for the
closest transits to have shallower absorption. As with Figure 9,
the likely distances for poorly constrained orbits are towards the
low dtr ends of their distributions, so the true correlation may be
as tight as suggested by the clouds of points for orbits derived
from accelerating features.
6 DISCUSSION AND CONCLUSIONS
This work is motivated by the realisation that previously esti-
mated orbital elements for β Pictoris' comets imply acceleration
of the absorption features should be visible in spectra taken over
several hours. Figures 2, 3 and 4 show that these accelerations
are indeed seen, and using a simple model these accelerating
features can be used to derive orbital elements for individual
comets. The constraints on the pericentre distance q and longi-
tude , and eccentricity e are degenerate, but constraints on the
distance to the star at the time of transit are less so. Thus, use-
ful constraints that can be directly compared to models of the
comets' origins can be derived in a model-independent way.
The constraints can be further improved with the reason-
able assumption that the orbits are parabolic, because comets
on low eccentricity orbits spend too much time near the star and
may not survive long enough to be observed. With this assump-
10
Grant M. Kennedy
tion, the pericentre distance and longitude are well constrained,
and the latter lies in the range 25 to 75◦ for the orbits modelled
here (Fig. 8). That is, the tendency for the pericentre longitudes
to lie in a restricted range suggested in the past based on orbits
estimated using an evaporation model (e.g. Beust et al. 1990) is
also found when the orbits are derived directly.
A further test of the evaporation model is of the predic-
tion that comet absorption is deeper for more distant transits,
because comae of closer comets are more strongly ablated and
therefore cover less of the stellar surface. This correlation is also
borne out by our analysis (Fig. 10).
Our technique of exocomet orbit fitting is analogous to
other techniques that measure the radial velocity of planets di-
rectly, by using cross-correlation with models of CO and H2O
absorption to show that their spectra are double-lined (e.g.
Snellen et al. 2010). Because they can measure the planet ve-
locity with a precision of a few percent, rather than the re-
flex motion of the host star, these techniques yield stringent
orbital constraints, for example on the eccentricity and inclina-
tion of non-transiting planets (Brogi et al. 2012). These obser-
vations have detected both absorption in the thermal emission
spectrum of the planet (e.g. Birkby et al. 2017), and absorption
in transmission spectra (i.e. during transit) (e.g. Snellen et al.
2010; Brogi et al. 2018). The latter detections are analogous to
the results presented here, and future transmission spectroscopy
of transiting planets may also yield independent and useful con-
straints on the orbital elements.
Our results are reasonably straightforward because the ap-
proach is simple. However, a few caveats and inbuilt assump-
tions should be noted. In fitting orbits, the stellar radius and
mass have been assumed fixed, so there remains a small sys-
tematic uncertainty in the results. We further assumed that the
transits cross the centre of the stellar disc, but there are no
cases where both ingress and egress are seen so this assump-
tion should not influence the results significantly. In some cases
the absorption lines are probably blended, meaning that a sin-
gle model may have been applied where two might have been
more appropriate; the most likely examples are the 80 km s−1
feature in the second panel and the two broad features between
50 and 150 km s−1 in the fifth panels in Figures 3 and 4. Again,
the results are unlikely to be significantly changed by use of a
more complex model, but the uncertainties on the derived orbital
parameters may increase. The most important issue for further
work to address is probably a systematic and objective identifi-
cation of accelerating features. Other issues include i) simulta-
neous fitting of H and K lines where possible, ii) simultaneous
fitting of possibly bended lines, and iii) modelling variable ab-
sorption during transit.
This work has various possible implications. Most obvi-
ously, the accelerating absorption features provide strong evi-
dence that the FEB phenomenon indeed originates in bodies that
pass in front of β Pictoris at close distances (for which comets is
the leading hypothesis). By fitting models to these features the
orbital properties are found to be consistent with previous esti-
mations that were based on an evaporation model, and therefore
these results largely validate that model as a reasonable physi-
cal explanation of the comet comae. Further work could attempt
to refine the evaporation model by attempting to again fit the
absorption features, but now using orbital element constraints
derived when acceleration is detected.
ACKNOWLEDGMENTS
GMK is supported by the Royal Society as a Royal Society
University Research Fellow. I am grateful to Matteo Brogi and
Daniel Bayliss for discussions during the preparation of this
work, and to the referee for a valuable review.
The code used in this research is available on github at
https://github.com/drgmk/feb-accel, including the final steps of
the chains used in Figures 6 to 10.
REFERENCES
Beust, H., Lagrange, A.-M., Plazy, F., & Mouillet, D. 1996,
A&A, 310, 181
Beust, H., Lagrange-Henri, A. M., Vidal-Madjar, A., & Ferlet,
R. 1989, A&A, 223, 304
Beust, H. & Lissauer, J. J. 1994, A&A, 282, 804
Beust, H. & Morbidelli, A. 1996, Icarus, 120, 358
-. 2000, Icarus, 143, 170
Beust, H., Vidal-Madjar, A., Ferlet, R., & Lagrange-Henri,
A. M. 1990, A&A, 236, 202
Birkby, J. L., de Kok, R. J., Brogi, M., Schwarz, H., & Snellen,
I. A. G. 2017, AJ, 153, 138
Blum, J., Gundlach, B., Krause, M., Fulle, M., Johansen,
A., Agarwal, J., von Borstel, I., Shi, X., Hu, X., Bentley,
M. S., Capaccioni, F., Colangeli, L., Della Corte, V., Fougere,
N., Green, S. F., Ivanovski, S., Mannel, T., Merouane, S.,
Migliorini, A., Rotundi, A., Schmied, R., & Snodgrass, C.
2017, MNRAS, 469, S755
Brogi, M., Giacobbe, P., Guilluy, G., de Kok, R. J., Sozzetti,
A., Mancini, L., & Bonomo, A. S. 2018, ArXiv e-prints
Brogi, M., Snellen, I. A. G., de Kok, R. J., Albrecht, S., Birkby,
J., & de Mooij, E. J. W. 2012, Nature, 486, 502
Crawford, I. A., Beust, H., & Lagrange, A.-M. 1998, MNRAS,
294, L31
Duncan, M., Quinn, T., & Tremaine, S. 1987, AJ, 94, 1330
Duncan, M. J. & Levison, H. F. 1997, Science, 276, 1670
Eiroa, C., Marshall, J. P., Mora, A., Montesinos, B., Absil, O.,
Augereau, J. C., Bayo, A., Bryden, G., Danchi, W., del Burgo,
C., Ertel, S., Fridlund, M., Heras, A. M., Krivov, A. V.,
Launhardt, R., Liseau, R., Lohne, T., Maldonado, J., Pilbratt,
G. L., Roberge, A., Rodmann, J., Sanz-Forcada, J., Solano,
E., Stapelfeldt, K., Th´ebault, P., Wolf, S., Ardila, D., Ar´evalo,
M., Beichmann, C., Faramaz, V., Gonz´alez-Garc´ıa, B. M.,
Guti´errez, R., Lebreton, J., Mart´ınez-Arn´aiz, R., Meeus, G.,
Montes, D., Olofsson, G., Su, K. Y. L., White, G. J., Barrado,
D., Fukagawa, M., Grun, E., Kamp, I., Lorente, R., Mor-
bidelli, A., Muller, S., Mutschke, H., Nakagawa, T., Ribas,
I., & Walker, H. 2013, A&A, 555, A11
Ferlet, R., Vidal-Madjar, A., & Hobbs, L. M. 1987, A&A, 185,
267
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J.
2013, PASP, 125, 306
Galland, F., Lagrange, A.-M., Udry, S., Chelli, A., Pepe, F.,
Beuzit, J.-L., & Mayor, M. 2006, A&A, 447, 355
Goodman, J. & Weare, J. 2010, Comm. App. Math and Comp.
Sci, 5, 65
Hobbs, L. M., Vidal-Madjar, A., Ferlet, R., Albert, C. E., &
Gry, C. 1985, ApJ, 293, L29
Kiefer, F., Lecavelier des Etangs, A., Boissier, J., Vidal-
c(cid:13) 0000 RAS, MNRAS 000, 000–000
Exocomet Orbit Fitting
11
Madjar, A., Beust, H., Lagrange, A.-M., H´ebrard, G., & Fer-
let, R. 2014, Nature, 514, 462
Koen, C., Balona, L. A., Khadaroo, K., Lane, I., Prinsloo, A.,
Smith, B., & Laney, C. D. 2003, MNRAS, 344, 1250
Kondo, Y. & Bruhweiler, F. C. 1985, ApJ, 291, L1
Lagrange, A.-M., De Bondt, K., Meunier, N., Sterzik, M.,
Beust, H., & Galland, F. 2012, A&A, 542, A18
Lagrange, A. M., Ferlet, R., & Vidal-Madjar, A. 1987, A&A,
173, 289
Lagrange, A.-M., Plazy, F., Beust, H., Mouillet, D., Deleuil,
M., Ferlet, R., Spyromilio, J., Vidal-Madjar, A., Tobin, W.,
Hearnshaw, J. B., Clark, M., & Thomas, K. W. 1996, A&A,
310, 547
Lagrange-Henri, A. M., Gosset, E., Beust, H., Ferlet, R., &
Vidal-Madjar, A. 1992, A&A, 264, 637
Lagrange-Henri, A. M., Vidal-Madjar, A., & Ferlet, R. 1988,
A&A, 190, 275
Levison, H. F., Duncan, M. J., & Wetherill, G. W. 1994, Na-
ture, 372, 441
Murray, C. D. & Dermott, S. F. 2000, Solar System Dynamics
Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013, Pro-
ceedings of the National Academy of Science, 110, 19273
Petterson, O. K. L. & Tobin, W. 1999, MNRAS, 304, 733
Pichierri, G., Morbidelli, A., & Lai, D. 2017, A&A, 605, A23
Quillen, A. C. & Holman, M. 2000, AJ, 119, 397
Sibthorpe, B., Kennedy, G. M., Wyatt, M. C., Lestrade, J.-F.,
Greaves, J. S., Matthews, B. C., & Duchene, G. 2018, MN-
RAS, 475, 3046
Smith, B. A. & Terrile, R. J. 1984, Science, 226, 1421
Snellen, I. A. G., de Kok, R. J., de Mooij, E. J. W., & Albrecht,
S. 2010, Nature, 465, 1049
Th´ebault, P. & Beust, H. 2001, A&A, 376, 621
c(cid:13) 0000 RAS, MNRAS 000, 000–000
12
Grant M. Kennedy
Table 1. Median parameters derived for accelerating features assuming parabolic orbits, and their standard deviations. MJD refers to the modified
Julian date on which the spectral observations began, and the table is listed in the same order as Figure 3.
MJD
d
σ vr
vr
km s−1 h−1
σvr
vr
km s−1
q
σq
σ
dtr
σdtr
∆
σ∆
R⋆
rad
R⋆
%
54542
54829
54829
54913
55170
55567
55597
56685
56694
56982
56988
57344
9
2
2
38
1
22
65
3
1
5
13
21
0.4
0.3
0.3
0.7
0.1
1.1
6.0
0.4
0.3
1.0
1.3
3.4
68
88
39
90
73
68
117
86
52
73
125
89
2.0
1.8
1.1
0.6
1.2
0.5
3.3
2.8
1.4
4.1
7.3
9.7
8
12
20
4
18
5
3
11
20
10
5
5
0.2
0.5
2
0.04
0.5
0.1
0.1
0.5
2
1
0.4
0.5
0.6
1.2
0.6
0.6
1.3
0.5
0.7
1.1
0.8
0.8
1.1
0.7
0.02
0.04
0.03
0.004
0.03
0.008
0.02
0.05
0.06
0.06
0.06
0.07
9
18
22
4
28
6
3
15
23
12
7
6
0.2
1
2
0.04
1
0.1
0.2
1
3
1
0.4
0.5
0.03
0.04
0.05
0.03
0.2
0.02
0.02
0.04
0.06
0.06
0.02
0.02
0.001
0.0009
0.001
0.0005
0.002
0.0008
0.001
0.002
0.002
0.003
0.002
0.002
c(cid:13) 0000 RAS, MNRAS 000, 000–000
|
1805.06918 | 1 | 1805 | 2018-05-17T18:30:09 | C/2016 R2 (PANSTARRS): A comet rich in CO and depleted in HCN | [
"astro-ph.EP"
] | We observed comet C/2016 R2 (PANSTARRS) with the ARO 10-m SMT, and report the first detection of CO emission from this comet with amounts high enough to be the primary driver of activity. We obtained spectra and maps of the CO J=2-1 rotational line at 230 GHz between 2017 December and 2018 January. We calculated an average production rate of Q(CO)=(4.6+/-0.4)x10$^{28}$ mol s$^{-1}$ at r ~2.9 au and delta ~2.1 au. The CO line is thin FWHM ~ 0.8 km s$^{-1}$ with a slight blue-shift ~ -0.1 km s$^{-1}$ from the ephemeris velocity, and we derive a gas expansion velocity of V$_{exp}$ = 0.50+/-0.15 km s$^{-1}$. This comet produced approximately half the CO that comet C/1995 O1 (Hale-Bopp) did at 3 au. If CO production scales with nucleus surface area, then the radius need not exceed ~15 km. The spectra and mapping data are consistent with CO arising from a combination of a sunward-side active area and an isotropic source. For HCN, we calculated a 3-sigma upper limit production rate of Q(HCN) < 8x10$^{24}$ molecules s$^{-1}$, which corresponds to an extraordinarily high abundance ratio limit of Q(CO)/Q(HCN) > 5000. We inferred a production rate of molecular nitrogen of Q(N$_2$) ~2.8x10$^{27}$ molecules s$^{-1}$ using our CO data and the reported N$_2$/CO column density ratio (Cochran & MacKay 2018a,b). The comet does not show the typical nitrogen depletion seen in comets, and the CO-rich, N$_2$-rich and HCN-depleted values are consistent with formation in a cold environment of T < 50 K that may have provided significant N$_2$ shielding. | astro-ph.EP | astro-ph |
Draft version May 21, 2018
Typeset using LATEX modern style in AASTeX62
C/2016 R2 (Pan-STARRS): A comet rich in CO and depleted in HCN
K. Wierzchos1 and M. Womack1
1University of South Florida
Department of Physics
4202 E. Fowler Ave
Tampa, FL, 33620 USA
ABSTRACT
We observed comet C/2016 R2 (Pan-STARRS) with the ARO 10-m SMT, and
report the first detection of CO emission from this comet with amounts high enough
to be the primary driver of activity. We obtained spectra and maps of the CO J=2-1
rotational line at 230 GHz between 2017 December and 2018 January. We calculated
an average production rate of Q(CO)=(4.6±0.4)x1028 mol s−1 at r ∼ 2.9 au and ∆ ∼
2.1 au. The CO line is thin (∆VF W HM ∼ 0.8 km s−1) with a slight blue-shift (δv ∼ -0.1
km s−1) from the ephemeris velocity, and we derive a gas expansion velocity of vexp
= 0.50 ± 0.15 km s−1. This comet produced approximately half the CO that comet
C/1995 O1 (Hale-Bopp) did at 3 au. If CO production scales with nucleus surface
area, then the radius need not exceed RR2 ∼ 15 km. The spectra and mapping data
are consistent with CO arising from a combination of a sunward-side active area and
an isotropic source. For HCN, we calculated a 3-sigma upper limit production rate
of Q(HCN) < 8x1024 molecules s−1, which corresponds to an extraordinarily high
abundance ratio limit of Q(CO)/Q(HCN) > 5000. We inferred a production rate of
molecular nitrogen of Q(N2) ∼ 2.8x1027 molecules s−1 using our CO data and the
reported N2/CO column density ratio (Cochran & McKay 2018a,b). The comet does
not show the typical nitrogen depletion seen in comets. The CO-rich, N2-rich and
HCN-depleted values are consistent with formation in a cold environment of T < 50
K that may have provided significant N2 shielding.
Keywords: comets, individual: C/2016 R2 (PanSTARRS)- solar system - astro-
chemistry - protoplanetary disks
Corresponding author: K. Wierzchos
[email protected]
2
Wierzchos and Womack
1. INTRODUCTION
Comets, comprised largely of ice and dust, constitute the least processed bodies of
the Solar System, and most travel around the Sun in eccentric orbits. Their nuclei
contain well-preserved samples of grains and gas from the protosolar nebula cloud in
which they formed Mumma & Charnley (2011). As a comet approaches the Sun, it
forms a coma around the nucleus by sublimating volatile ices, which release dust and
other gases. In order to constrain models of comet (and by extension solar system)
formation it is vital to accurately determine the composition and physical state of
cometary nuclei Combi & Fink (1997); Cochran et al. (1999).
Comet C/2016 R2 (Pan-STARRS) -hereafter R2- was discovered at r = 6.3 au from
the Sun on 2016 September 7 when it exhibited a 20′′ wide coma at 19.1 visible
magnitude (Weryk & Wainscoat 2016). It has an estimated orbital period of 20,000
years, a highly eccentric orbit tilted at an angle of 58 deg to the ecliptic, and a semi-
major axis of a ∼ 740 au. These orbital characteristics identify the object as an Oort
Cloud comet, but not dynamically new, since it has presumably already had many
journeys through the inner solar system (Levison 1996).
Upon discovery, this comet exhibited a coma at a distance where most comets
appear inactive. Water is the dominant ice in all comets and is not heated enough
by the Sun to sublimate efficiently until much closer, typically r = 2 − 3 au. Thus,
comet R2 also receives the "distantly active comet" classification. Instead of water-
ice sublimation, the observed comae of distant comets are generally considered to be
due to release of cosmogonically abundant hypervolatile species, such as CO and/or
CO2 (Ootsubo et al. 2012; Reach et al. 2013; Bauer et al. 2015; Womack et al. 2017;
Wierzchos et al. 2017).
By late 2017, optical images of R2 revealed a deep-blue-colored coma and ion-tail
with an absence of dust. The blue color in the coma is largely due to emission from
CO+ (a photoionization production of CO) and to some extent, N+
2 , which were both
observed to be strong, with a ratio of N2/CO = 0.06, one of the highest ever reported
for a comet (Cochran & McKay 2018a,b). The strong presence of CO+, and the lack
of CO+
2 emission in these optical spectra indicate that the comet's activity is probably
dominated by the outgassing properties of CO and not CO2. The comet's high N+
2
abundance is very important, because its likely parent, N2 is typically depleted in
comets, and like CO, N2 sublimates at extremely low temperatures (Iro et al. 2003;
Womack et al. 2017). The abundance of N2 is high enough that it may play a substan-
tial role in R2's distant outgassing behavior. Furthermore, N2 is also an important
molecule for astrochemical models of the solar system and other planetary systems
(Fegley & Prinn 1989; Lodders & Fegley 2010; Moses et al. 2016).
2. OBSERVATIONS AND RESULTS
We used the Arizona Radio Observatory 10-m Submillimeter Telescope in order to
search for CO J=2-1 (at 230.53799 GHz) and HCN J=3-2 (at 265.88643 GHz) emission
C/2016 R2: A comet rich in CO and depleted in HCN
3
in comet C/2016 R2 during 2017 December - 2018 January when its heliocentric and
geocentric distances were r ∼ 2.9 au and ∆ ∼ 2.1 au, respectively (Table 1).
Table 1: Observations of Comet C/2016 R2
Molecule
UT Date1
r (au) ∆(au) T∗
Adv (K km s−1) Q (x1028 mol s−1)
CO(2-1)
2017-12-22.21
2017-12-23.09
2017-12-30.14
2017-12-31.13
2018-01-16.03
HCN(3-2) 12-23.19 & 01-16.15
2.98
2.97
2.94
2.93
2.86
2.92
2.05
2.05
2.06
2.06
2.15
2.10
0.26±0.01
0.28±0.02
0.26±0.02
0.27±0.02
0.27±0.02
<0.0302
4.4±0.2
4.6±0.4
4.6±0.4
4.6±0.4
4.6±0.4
< 0.0008
1Dates listed represent the midpoint of data collection.
2The T*Adv line area upper limit for HCN was calculated using three times the rms of
the HCN spectrum multiplied by an assumed linewidth of 1 km s−1. For HCN, r and ∆
are the average of the comet's distances on December 23, 2017 and January 16, 2018.
We used the dual polarization 1.3 mm receiver with ALMA Band 6 sideband-
separating mixers for all observations. The mode of data acquisition was beam-
switching mode with a reference position of +2′ in azimuth. An integration time of
3 minutes on the source and 3 minutes on the sky reference position for each scan
was used. System temperatures were typically in the low 300 K for all the data. The
temperature scale for all SMT receiver systems, T∗
A, was determined by the chopper
wheel method, with TR=T∗
A/ηb, where TR is the temperature corrected for beam
efficiency and ηb is the main beam efficiency of the SMT with a value of ηb = 0.74
for both the CO and HCN frequencies. The backends consisted of a 2048 channel 1
MHz filterbank used in parallel (2 x 1024) mode and a 250 kHz/channel filterbank
also in parallel (2 x 250). The 250 kHz/channel filterbanks provided the equivalent
velocity resolutions of 0.325 km s−1 for CO J=2-1 and 0.282 km s−1 for HCN J=3-2.
The 1 MHz resolution filterbanks were significantly broader than the expected CO
and HCN linewidths and thus were not used in the analysis. The ARO SMT beam
size is θB = 32.7′′ at the CO J=2-1 frequency and 28.4′′ at the HCN J=3-2 frequency.
After every six scans on the comet, we updated the pointing and focus on Uranus
and on the strong radio-source Orion-A. The pointing and tracking were showing an
accuracy of < 1′′ RMS throughout the observing epochs. The comet's phase angle
ranged from 7 degrees (on 2017 Dec 22) to 15 degrees (on 2018 Jan 16). CO emission
was detected and remained relatively constant during this time (Table 1).
The CO J=2-1 line was detected in R2 during a single six minute scan on the
UT 2017 Dec 22 observations, and the total for the first day is shown in Fig-
ure 1. The spectrum in Figure 1 is the first detection of CO emission in this
comet, which we announced to the astronomical community in a preliminary report
4
Wierzchos and Womack
CO(2-1) Spectrum of C/2016 R2 PANSTARRS
0.3
0.2
0.1
0.0
)
K
(
A
*
T
0.33 km s-1 / channel
2017 Dec 22.21 UT
r = 2.98 au
-20
-15
-10
-5
0
5
10
15
20
Offset Velocity (km s-1)
Figure 1. The first detection of CO emission in C/2016 R2 on UT 2017 Dec 22, when the
comet was at r = 2.9 and ∆ = 2.1 au. The CO J=2-1 rotational line was bright with an
intensity of TA* = 0.3 K easily detected in single scans with the Arizona Radio Observatory
Submillimeter 10-m telescope. The spectrum was obtained with 250 kHz/channel spectral
resolution with a Gaussian measured linewidth of FWHM = 0.79 km s−1 and was not
significantly shifted from the comet ephemeris velocity, which is indicated by a vertical
dashed line at zero offset velocity.
(Wierzchos & Womack 2017). Both polarizations showed the line, and little change
was observed in the line intensity, shape or area throughout the observing period.
The ∼ 3 channel wide (FWHM) line profile yielded a ∆VF W HM ∼ 0.79 ± 0.33 km
s−1 if a Gaussian fit is assumed. The line is typically blue-shifted by a small amount
ranging from δv = -0.08 to -0.18 km s−1.
We also mapped the CO emission on UT 2018 Jan 16.1 to assess its spatial extent
in the inner coma. The map was constructed with a 9-point grid technique centered
on the nucleus position. The map had 16′′ spacings and integrations of 6 minutes
for each position. The pointing separations for the map are equal to the half-power
beamwidth (HPBW) of the SMT 10-m dish at this frequency (see Figure 2). The
map was aligned along the RA and Dec axis and the direction to the Sun is indicated
in the figure.
We searched for the HCN J=3-2 transition on 2017 Dec 23 and 2018 Jan 16 and
did not detect a line down to a cumulative 1-sigma level of TA* = 0.010 K in the 250
kHz/channel filterbanks (Table 1). We also searched for, and did not find, CH3OH
(251 GHz), H2CO (218 GHz), N2H+ (279 GHz), HCO+ (276 GHz) and CS (244 GHz).
The significance of those non-detections and limits will be discussed in a later paper.
C/2016 R2: A comet rich in CO and depleted in HCN
5
CO J=2-1 emission in comet C/2016 R2
)
K
(
*
A
T
Offset velocity (km s-1)
Figure 2. Map of CO emission from C/2016 R2 constructed on 2018 January 16 with
the ARO SMT. The size of the map corresponds to 96,000 km x 96,000 km (64′′x 64′′) on
the sky at the comet's projected distance. The direction to the Sun is toward the right as
indicated in the Figure. The comet's ephemeris speed is indicated with vertical dashed lines
at zero velocity. CO emission peaked in intensity at the ephemeris location of the nucleus
(center position) and may be slightly increased on the sunward side when compared to the
tailward side.
3. ANALYSIS AND DISCUSSION
3.1. CO outgassing velocities and spatial extent
Important modeling parameters, such as expansion velocity, vexp, and outgassing
patterns, can be extracted from spectral line profiles and maps that have sufficiently
high resolution (Biver et al. 2002). Comet activity models typically consider two
different sources for CO in comets, one emanating from the subsolar point where
solar heating is greatest, and one from an isotropic source in the coma. Here we
briefly address how the R2 CO data aligns with the models.
First, we examine the CO line profile, which has a single velocity component and
is slightly blue-shifted from the comet's ephemeris velocity by δv = -0.12 ± 0.20 km
s−1 (see Figure 1). The line is between 2-3 channels wide (corresponding to FWHM
of 0.66 - 0.99 km s−1), and by fitting a Gaussian, we derived a FWHM linewidth of
∆vF W HM = 0.85 ± 0.33 km s−1. Thus, the data are consistent with having a FWHM
linewidth of ∼ 0.8 km s−1. The small velocity shift we observe of ∼ -0.1 km s−1 for the
comet with a low phase angle is consistent with at least some of the outgassing takes
place on the Earth-facing side. The half-width half-maximum (HWHM) linewidth
6
Wierzchos and Womack
measured on the blue-ward wing is proportional to the outflow velocity of the gas (see
Biver et al. (1999)), and thus, we estimate the gas expansion velocity to be vexp = 0.50
± 0.15 km s−1. This is comparable to what was measured in other comets at this same
heliocentric distance, such as C/1995 O1 (Hale-Bopp) and C/2006 W3 (Christensen)
(Gunnarsson et al. 2003; Bockel´ee-Morvan et al. 2010). de Val-Borro et al. (2018)
also report seeing CO emission from this comet approximately three weeks after our
first detection, but with a velocity redshift of ∼ +1 km s−1 and a somewhat broader
linewidth of 1.0-1.3 km s−1. We also observed the comet during this time and do not
confirm their redshifted velocity component.
The nearly-centered and narrow CO spectrum of R2 is consistent with a simple
model of cool gas expanding isotropically with vexp ∼ 0.50 km s−1, which is what
we used to calculate production rates. The data are also consistent with a more
detailed scenario if one looks more carefully at the spectral line profile. First, we
revisit Hale-Bopp and Christensen, which also produced substantial CO at r = 3 –
4 au. The CO emission in these comets were also slightly offset from the ephemeris
velocity by δv ∼ -0.1 km s−1 (Biver 1997; Womack et al. 1997; Bockel´ee-Morvan et al.
2010). For Hale-Bopp, the CO line profiles were fit by a detailed two-component
model comprised of isotropic outgassing of cold CO gas, combined with a blue-shifted
velocity component associated with a sunward side active area (Gunnarsson et al.
2003). This model produces two peaked lines in comets beyond 4 or 5 au and single
peaked lines for comets within 4 au. Therefore, the CO emission is also consistent
with CO arising from a mix of a sunward-side active area and a symmetric source
either in the coma or from the nucleus.
We also examined the maps for clues about the CO outgassing. Figure 2 shows
that the emission peaks at the nucleus position provided by the ephemeris. Further-
more, CO emission was readily detected at all positions out to at least 45′′ with a
decrease in intensity by 20-40% relative to the line at the center position, consistent
with isotropic outflow of CO. There is evidence for a slight sunward enhancement
of emission on the sunward side compared to the tailward side. This gives further
support for contribution from an active area releasing CO on the sunward side, as is
also indicated from the spectral line profile.
It is not clear what could generate isotropic CO emission in the coma. Given that
CO+
2 was not detected in the optical spectrum of R2 (Cochran & McKay 2018a) and
we did not see CH3OH or H2CO down to significant limits (Wierzchos, in prep.), it is
not likely that CO was produced in significant amounts by photodissociation of these
species, which are plausible secondary sources for CO in other comets. Also, this
comet has an almost nonexistent dust coma, and thus CO is not likely to come from
refractory cometary grains in the coma. Perhaps additional CO is released by subli-
mating water ice grains in the coma that were ejected from the nuclear sunward-side
facing CO source. Measurements of OH or H2O emission would be useful constraints
C/2016 R2: A comet rich in CO and depleted in HCN
7
to the CO production model in R2. Much higher resolution spectra, ≤ 0.1 km s−1,
would also be valuable for testing models of CO production.
In principle, more detailed modeling of the spectral line profile and mapping data
could significantly constrain models of the release mechanisms for CO and physical
conditions in the coma, but this requires higher spectral and spatial resolution data,
and is beyond the scope of this paper.
3.2. Production rates of CO and HCN
We calculated column densities assuming the excitation was dominated by
following the modeling described in
collisional and fluorescence contributions,
Crovisier & Le Bourlot (1983), Bockelee-Morvan & Crovisier
(1985), and Biver
(1997). We assumed a rotational and excitation temperature of 25K, which is consis-
tent with the empirical fit to Hale-Bopp CO data described in Biver et al. (2002).
The column density for CO was fairly constant, with an average value of N(CO) =
(1.89 ± 0.14)x1014 cm−2. In order to calculate production rates, we assumed a gas
expansion velocity of 0.50 km s−1, which is consistent with the CO spectral line profile
and values from other comets at this distance, as described in Section 3.1. Using a
photodissociation decay model (Haser 1957) and assuming isotropic outgassing of CO
we find an average production rate of Q(CO) = (4.6 ± 0.4)x1028 mol s−1 between
2017 December 22 and 2018 January 16 (see Table 1). de Val-Borro et al. (2018)
report higher CO production rates, despite reporting similar line intensities. We think
the different production rates are the result of using different modeling parameters,
such as expansion velocity. There is insufficient detail about the modeling in their
preliminary announcement to warrant further comments.
As discussed in section 3.1, CO2 is not likely to be a significant parent of CO in
this comet.
Infrared observations of CO2 would be very useful in quantifying any
contributions from CO2 to the coma and/or CO emission. Also, to date, no searches
for OH or H2O emission have succeeded, which implies that water-ice sublimation is
probably not responsible for most of R2's activity. Thus, the major driver at this
distance is probably CO outgassing.
The CO production rate of R2 is very high and approximately half that of C/1995
O1 Hale-Bopp at this same distance from the Sun. Based on the high Q(CO) values,
we consider R2 to be "CO-rich." Other CO-rich comets, typically have CO/H2O >
8% (Dello Russo et al. 2016) and the ratio for R2 may be substantially higher than
8%, since water has yet to be detected. We point out that these observations were
obtained when the comet was at ∼ 3 au from the Sun, which is too far for water-ice
to sublimate efficiently. Thus, the relative CO/H2O content in the nucleus may be
much higher than we can determine at this heliocentric distance.
The nucleus' radius, RR2, can be estimated based on the assumption that Q(CO) is
proportional to the nucleus surface area and the insolation received, and then com-
pared to comets at the same heliocentric distance for which both Q(CO) and radius
8
Wierzchos and Womack
are independently known. For example, if we use RHB = 30 km (Fern´andez 2002)
and Q(CO) = 1.8x1029 mol s−1 for Hale-Bopp (at 3 au) (Biver et al. 2002), then the
average Q(CO) = 4.6 x1028 mol s−1 at ∼ 3 au corresponds to a radius of R ∼ 15 km.
Similarly, from a comparison with comet Christensen's radius upper limit of RCh <
13 km (Korsun et al. 2016) and Q(CO) = 3.9x1028 mol s−1 (Bockel´ee-Morvan et al.
2010) at 3 au, we derived that R < 14 km. CO activity may not, in fact, scale directly
with surface area for this comet, but if it does, then we find that the nucleus radius
need not exceed RR2 ∼ 15 km in order to explain the measured CO production rate.
We derived a 3-sigma upper limit of Q(HCN) < 8.0x1024 molecules s−1 from all
the HCN data. For comparison, this is ∼ 100 times lower than observed for Hale-
Bopp at the same distance (Biver et al. 2002). Our non-detection of HCN emission
is consistent with the absence of the CN band at 3880 Angstroms, as reported by
Cochran & McKay (2018a), assuming that CN is caused by photolysis of HCN.
Table 2: Compiled Q(CO)/Q(HCN) ratios in CO-rich and other comets
Reference
Q(CO)/Q(HCN) r* (au)
Comet
C/2016 R2 (Pan-STARRS)
29P/Schwassmann-Wachmann 1
C/2006 W3 (Christensen)
C/1995 O1 (Hale-Bopp)
C/2010 G2 (Hill)
C/1996 B2 (Hyakutake)
C/1999 T1 (McNaught-Hartley)
C/2001 Q4 (NEAT)
C/2009 P1 (Garrad)
C/2013 R1 (Lovejoy)
Oort Cloud Comets
Jupiter Family Comets
All comets
>5000
3300†
243
125-650
52-91
70
2.9
5.8
3.2
3
0.9
2.5
96 0.6, 0.7
46
31
1.3
1.0
This paper
[1,20]
[21]
[3]
[2,3,7]
[16]
[8]
[9,11]
[14]
36 1.6, 2.1 [12,13,15,18]
34
28
9
25
1.3
-
-
-
[17]
[19]
[19]
[19]
[4] Disanti et al. (2002),
[7] Brooke et al. (2003),
References:
[2] Magee-Sauer et al. (1999),
[1] Cochran & Cochran (1991),
[3]
[5] Dello Russo et al. (2002, 2004),
Biver et al. (2002),
[6] Magee-Sauer et al. (2002),
[8] DiSanti et al. (2003),
[9] Gibb et al. (2003), [10] Kawakita et al. (2003), [11] Mumma et al. (2003), [12]
Paganini et al. (2012), [13] Villanueva et al. (2012), [14] de Val-Borro et al. (2013),
[15] DiSanti et al. (2014), [16] Kawakita et al. (2014), [17] Paganini et al. (2014), [18]
McKay et al. (2015), [19] Dello Russo et al. (2016), [20] Womack et al. (2017), [21]
Bockel´ee-Morvan et al. (2010)
C/2016 R2: A comet rich in CO and depleted in HCN
9
*r is the heliocentric distances at which production rates were measured. Two
values of r are listed when CO and HCN measurements were not simultaneous.
†We assumed Q(HCN) ∼ Q(CN) for 29P, see Womack et al. (2017).
We briefly compare the CO and HCN production rates with other comets, since
emission from these two species are commonly detected and their ratio may provide
insights to the chemical composition of the nucleus and/or coma (Table 2). The
comets identified by name are those reported to be CO-rich. Also listed in the table
are average values for larger groups of comets, such as Oort Cloud, Jupiter Family,
and finally an "all comets" average value. As the table shows, the relative production
rate value derived for R2 is extraordinarily high: Q(CO)/Q(HCN) > 5000 at r ∼ 3
au. The average value for all comets measured is Q(CO)/Q(HCN) ∼ 25 and this ratio
varies by less than a factor of three between Jupiter Family Comets (JFCs) and Oort
Cloud Comets (OCCs). The only group where it noticeably departs from the average
value is for CO-rich comets, which are listed individually in the top panel of the table.
It is perhaps not surprising that the comets designated as CO-rich also have elevated
Q(CO)/Q(HCN) values, but even among these CO-rich comets the limit derived for
R2 is the highest values to date for any comet.
The very high Q(CO) and very low Q(HCN) in R2 is difficult to understand in
terms of typical comet compositions. At 3 au, R2's comet nucleus has not received
much solar heating and so it will preferentially release CO over HCN, due to its higher
volatility. This behavior was seen in the abundance ratio of CO/HCN in Hale-Bopp,
which decreased as the comet got closer to the Sun and more HCN was released (see
Biver et al. (2002) and Table 2). There are not many measurements of both CO and
HCN in comets at ∼ 3 au, but R2's value is substantially higher than those measured
for the CO-rich comets Hale-Bopp, Christensen or C/2010 G2 Hill in the range of
2.5 - 3.0 au, suggesting that R2's high CO/HCN ratio cannot be explained solely
due to volatility differences between the two molecules.
Interestingly, the highest
CO/HCN values were obtained in comets known to be both distantly active and
CO-rich (R2, 29P, Christensen and Hale-Bopp). This is worth looking into further,
but the data are sparse. Even for a comet at 3 au, the HCN upper limit that we
derived is extraordinarily low. Another possible clue is that R2's coma is largely
gaseous with very little dust (Cochran & McKay 2018a), and this may be related
to the significantly decreased amounts of HCN and other volatiles. The chemical
composition of R2's coma is noticeably atypical when compared to other comets.
3.3. High N2 Production Rates
Searching for additional clues to the unusual chemical composition of this comet,
we now turn our attention to molecular nitrogen. Measuring cometary N2 is of con-
siderable importance for many reasons, including testing models of the condensation
and incorporation of ices in the protosolar nebula, and calculating the N2/NH3 abun-
dance ratio, which is a key diagnostic of primordial physical and chemical conditions
10
Wierzchos and Womack
(Fegley & Prinn 1989; Womack et al. 1992). N2 is also a highly volatile molecule
and it can contribute to comet activity if significantly incorporated in the nucleus.
Furthermore, N2 is trapped and released in a manner similar to Argon, and thus, de-
tecting N+
2 emission in any coma suggests that Ar may also be present in high amounts
(Owen & Bar-Nun 1995). Despite its importance, it has been difficult to measure the
N2 abundances for all but a few comets (Lutz et al. 1993; Cochran 2002; Korsun et al.
2014; Rubin et al. 2015; Ivanova et al. 2016). Strikingly, N+
2 optical emission is
clearly detected in R2, with a measured abundance ratio of N(N2)/N(CO)=0.06
(Cochran & McKay 2018a,b).
The N2/CO ratio is an important observational constraint for testing the formation
environment for cometary ice; it is determined by the temperature of the gases when
they were incorporated into the ice as well as any subsequent processing that may have
preferentially affected one volatile over the other. In order to place this in context,
we briefly review two general scenarios for comet formation: one where comets ag-
glomerated from pristine amorphous water ice grains originating from the interstellar
medium onto which N2 and CO condensed in the protosolar nebula (Owen & Bar-Nun
1993). This model proposes that the N2/CO ratio in ices strongly depends on the
temperature of the materials at the time the volatiles condensed or were trapped.
The ratio derived from R2's optical spectra of N+
2 and CO+ agrees very well with the
predicted value of N2/CO = 0.06 for icy planetesimals forming in the solar nebula at
about 50 K (Owen & Bar-Nun 1995; Iro et al. 2003). Alternately, comets may have
agglomerated from crystalline water-ice grain clathrates that trap N2 and CO. Due
to its relatively small size, N2 is not readily trapped by clathrates, which leads to a
lower predicted ratio of N2/CO ranging from ∼ 0.002 to 0.02 (Mousis et al. 2012).
Thus, perhaps the measured value of N2/CO = 0.06 in R2 is not consistent with a
clathrate model.
In addition to being relatively rich in N2, we note that this comet is severely depleted
in HCN, as discussed in the previous section. Physicochemical models of nitrogen
chemistry in protostellar disks show that photodissociation of N2 leads to production
of HCN (Hily-Blant et al. 2017). It is interesting to consider that comet R2 may have
formed in the protosolar nebula disk where there was significant N2 shielding that led
to the high N2 and decreased HCN abundances.
N2 cannot undergo rotational transitions due to lack of a permanent dipole moment,
and thus emits no radiation at millimeter-wavelengths. However, because the N2
abundance can put such strong constraints on comet formation models, we derived an
N2 column density and production rate using the N2/CO abundance ratio calculated
from optical spectra and our CO results. We chose the CO data from 2017 Dec 22,
because it is closest in time to the Cochran & McKay (2018b) N2/CO value on 2017
Dec 8-10. We derived an N2 column density of N(N2) = (1.1 ± 0.2)x1013 cm−2 and
a production rate of Q(N2) = (2.8 ± 0.4)x1027 molecules s−1. This production rate
corresponds to a mass loss rate of 130 kg s−1 for N2. Determining the N2 production
C/2016 R2: A comet rich in CO and depleted in HCN
11
rate in this manner can be useful in comparison with those of water and/or NH3,
if these volatiles' abundances are established through direct observation or via their
daughter products (Tegler et al. 1992; Crovisier 1989).
3.4. Conclusions
We report the first detection of neutral CO emission, and an upper limit of HCN,
in comet C/2016 R2 Pan-STARRS at r ∼ 3 au. The CO line profile shape is charac-
teristic of CO emission seen in other comets at this distance, with a linewidth of ∼
0.8 km s−1 and is slightly blue-shifted from the ephemeris velocity. A 64′′x 64′′ map
(∼ 96,000 km x 96,000 km at the comet's projected distance) shows that the CO
emission peaks in intensity at the ephemeris position and decreases by 20-40% at the
off-centered positions. The spectra and map are consistent with CO arising from a
combination of an isotropic source and an active area on the sunward side.
If comet R2's CO output is proportional to surface area of the nucleus, then we
find that the radius need not exceed RR2 ∼ 15 km in order to explain the measured
CO production rate. Thus, R2 may be larger-than-average in size, but need not be a
giant comet in order to explain the measured CO production rates.
The very large amount of CO, and the apparent absence, or very low outgassing
rate, of HCN, leads to a CO/HCN production rate ratio over 5000. This is remarkably
high compared to other comets at r = 3 au, even when including other comets known
to be CO-rich. The high CO/HCN ratio cannot be explained solely due to volatility
differences between the two, and may represent a compositional difference between
R2 and most other comets. When considered along with the high outgassing rate of
N2, it is interesting to consider whether comet R2 formed in a region of the protosolar
nebula with substantial N2 shielding, which could have led to higher N2 and decreased
HCN abundances.
N2 production rates were derived from the N2/CO ratio (Cochran & McKay 2018b)
and our CO production rates, and were calculated to be Q(N2) = (2.8 ± 0.4)x1027
molecules s−1 at 3 au. N2 production rates will be valuable for comparison with those
of water and/or NH3, if detected in this comet.
R2's coma composition is clearly very different from other comets observed thus far,
both in the high N2 abundance, and significant decrease in other typically abundant
molecules, such as HCN. Further observations of this comet along all spectral ranges
are highly encouraged, especially those of NH2 in the optical, and NH3, CO2 and
CH4 in the infrared in order to measure the key diagnostic ratios N2/NH3, CO/CO2
and CO/CH4, which will provide observational tests for formation models of the
environment of this comet.
M.W. and K.W. acknowledge support from NSF grant AST-1615917. We also
thank the ARO 10m SMT crew. The SMT is operated by the ARO, the Steward
Observatory, and the University of Arizona, with support through the NSF University
12
Wierzchos and Womack
Radio Observatories program grant AST-1140030. This work was completed with the
GLIDASS CLASS software: http://www.iram.fr/IRAMFR/GILDAS
Facilities: Arizona Radio Observatory (SMT)
C/2016 R2: A comet rich in CO and depleted in HCN
13
Bauer, J. M., Stevenson, R., Kramer, E.,
Disanti, M. A., dello Russo, N.,
REFERENCES
et al. 2015, ApJ, 814, 85
Biver. 1997, Universit Paris 7, PhDT
Biver, N., Bockel´ee-Morvan, D., Crovisier,
J., et al. 1999, AJ, 118, 1850
Biver, N., Bockel´ee-Morvan, D., Colom,
P., et al. 2002, EM&P, 90, 5
Bockelee-Morvan, D., & Crovisier, J.
1985, A&A, 151, 90
Bockel´ee-Morvan, D., Hartogh, P.,
Crovisier, J., et al. 2010, A&A, 518,
L149
Brooke, T. Y., Weaver, H. A., Chin, G.,
et al. 2003, Icarus, 166, 167
Cochran, A. L. 2002, ApJL, 576, L165
Cochran, A. L., Barker, E. S., & Cochran,
W. D. 1999, in B.A.A.S., Vol. 31, #190,
1101
Cochran, A. L., & Cochran, W. D. 1991,
Icarus, 90, 172
Cochran, A. L., & McKay, A. J. 2018a,
ArXiv e-prints, arXiv:1801.01199
-. 2018b, ApJL, 856, L20
Combi, M. R., & Fink, U. 1997, Astroph.
J., 484, 879
Crovisier, J. 1989, A&A, 213, 459
Crovisier, J., & Le Bourlot, J. 1983,
A&A, 123, 61
de Val-Borro, M., Milam, S. N., Cordiner,
M. A., et al. 2018, The Astronomer's
Telegram, 11254
de Val-Borro, M., Kuppers, M., Hartogh,
P., et al. 2013, A&A, 559, A48
Dello Russo, N., Disanti, M. A.,
Magee-Sauer, K., Gibb, E., & Mumma,
M. J. 2002, in ESA Special Publication,
Vol. 500, Asteroids, Comets, and
Meteors: ACM 2002, ed. B. Warmbein,
689–692
Dello Russo, N., DiSanti, M. A.,
Magee-Sauer, K., et al. 2004, Icarus,
168, 186
Dello Russo, N., Kawakita, H., Vervack,
R. J., & Weaver, H. A. 2016, Icarus,
278, 301
Magee-Sauer, K., et al. 2002, in ESA
Special Publication, Vol. 500,
Asteroids, Comets, and Meteors: ACM
2002, ed. B. Warmbein, 571–574
DiSanti, M. A., Mumma, M. J., Dello
Russo, N., Magee-Sauer, K., & Griep,
D. M. 2003, Journal of Geophysical
Research (Planets), 108, 5061
DiSanti, M. A., Villanueva, G. L.,
Paganini, L., et al. 2014, Icarus, 228,
167
Fegley, Jr., B., & Prinn, R. G. 1989, in
The Formation and Evolution of
Planetary Systems, ed. H. A. Weaver &
L. Danly, 171–205
Fern´andez, Y. R. 2002, Earth Moon and
Planets, 89, 3
Gibb, E. L., Mumma, M. J., Dello Russo,
N., DiSanti, M. A., & Magee-Sauer, K.
2003, Icarus, 165, 391
Gunnarsson, M., Bockel´ee-Morvan, D.,
Winnberg, A., et al. 2003, A&A, 402,
383
Haser, L. 1957, BSRSL, 43, 740
Hily-Blant, P., Magalhaes, V., Kastner,
J., et al. 2017, A&A, 603, L6
Iro, N., Gautier, D., Hersant, F.,
Bockel´ee-Morvan, D., & Lunine, J. I.
2003, Icarus, 161, 511
Ivanova, O. V., Luk'yanyk, I. V., Kiselev,
N. N., et al. 2016, P&SS, 121, 10
Kawakita, H., Watanabe, J.-i., Kinoshita,
D., Ishiguro, M., & Nakamura, R. 2003,
ApJ, 590, 573
Kawakita, H., Dello Russo, N., Vervack,
Jr., R., et al. 2014, ApJ, 788, 110
Korsun, P. P., Kulyk, I., Ivanova, O. V.,
et al. 2016, A&A, 596, A48
Korsun, P. P., Rousselot, P., Kulyk, I. V.,
Afanasiev, V. L., & Ivanova, O. V.
2014, Icarus, 232, 88
Levison, H. F. 1996, in Astronomical
Society of the Pacific Conference Series,
Vol. 107, Completing the Inventory of
the Solar System, ed. T. Rettig & J. M.
Hahn, 173–191
Lodders, K., & Fegley, B. 2010,
Chemistry of the Solar System
14
Wierzchos and Womack
Lutz, B. L., Womack, M., & Wagner,
Paganini, L., Mumma, M. J., Villanueva,
R. M. 1993, ApJ, 407, 402
Magee-Sauer, K., dello Russo, N., Disanti,
M. A., Gibb, E., & Mumma, M. J.
2002, in ESA Special Publication, Vol.
500, Asteroids, Comets, and Meteors:
ACM 2002, ed. B. Warmbein, 549–552
Magee-Sauer, K., Mumma, M. J.,
DiSanti, M. A., Russo, N. D., & Rettig,
T. W. 1999, Icarus, 142, 498
McKay, A. J., Cochran, A. L., DiSanti,
M. A., et al. 2015, Icarus, 250, 504
Moses, J. I., Marley, M. S., Zahnle, K.,
et al. 2016, ApJ, 829, 66
Mousis, O., Guilbert-Lepoutre, A.,
Lunine, J. I., et al. 2012, ApJ, 757, 146
Mumma, M. J., & Charnley, S. B. 2011,
ARA&A, 49, 471
Mumma, M. J., DiSanti, M. A., Dello
Russo, N., et al. 2003, Advances in
Space Research, 31, 2563
G. L., et al. 2012, ApJL, 748, L13
-. 2014, ApJ, 791, 122
Reach, W. T., Kelley, M. S., &
Vaubaillon, J. 2013, Icarus, 226, 777
Rubin, M., Altwegg, K., Balsiger, H.,
et al. 2015, Science, 348, 232
Tegler, S. C., Burke, L. F., Wyckoff, S.,
et al. 1992, ApJ, 384, 292
Villanueva, G. L., Mumma, M. J.,
DiSanti, M. A., et al. 2012, Icarus, 220,
291
Weryk, R., & Wainscoat, R. J. 2016,
Central Bureau Electronic Telegrams,
4318
Wierzchos, K., & Womack, M. 2017,
Central Bureau Electronic Telegrams,
4464
Wierzchos, K., Womack, M., & Sarid, G.
2017, AJ, 153, 230
Womack, M., Festou, M. C., & Stern,
S. A. 1997, AJ, 114, 2789
Ootsubo, T., Kawakita, H., Hamada, S.,
Womack, M., Sarid, G., & Wierzchos, K.
et al. 2012, ApJ, 752, 15
Owen, T., & Bar-Nun, A. 1993, Nature,
361, 693
-. 1995, Icarus, 116, 215
2017, PASP, 129, 031001
Womack, M., Wyckoff, S., & Ziurys,
L. M. 1992, ApJ, 401, 728
|
1605.02731 | 1 | 1605 | 2016-05-09T20:00:00 | Imaging Extrasolar Giant Planets | [
"astro-ph.EP"
] | High-contrast adaptive optics imaging is a powerful technique to probe the architectures of planetary systems from the outside-in and survey the atmospheres of self-luminous giant planets. Direct imaging has rapidly matured over the past decade and especially the last few years with the advent of high-order adaptive optics systems, dedicated planet-finding instruments with specialized coronagraphs, and innovative observing and post-processing strategies to suppress speckle noise. This review summarizes recent progress in high-contrast imaging with particular emphasis on observational results, discoveries near and below the deuterium-burning limit, and a practical overview of large-scale surveys and dedicated instruments. I conclude with a statistical meta-analysis of deep imaging surveys in the literature. Based on observations of 384 unique and single young ($\approx$5--300~Myr) stars spanning stellar masses between 0.1--3.0~\Msun, the overall occurrence rate of 5--13~\Mjup \ companions at orbital distances of 30--300~AU is 0.6$^{+0.7}_{-0.5}$\% assuming hot-start evolutionary models. The most massive giant planets regularly accessible to direct imaging are about as rare as hot Jupiters are around Sun-like stars. Dividing this sample into individual stellar mass bins does not reveal any statistically-significant trend in planet frequency with host mass: giant planets are found around 2.8$^{+3.7}_{-2.3}$\% of BA stars, $<$4.1\% of FGK stars, and $<$3.9\% of M dwarfs. Looking forward, extreme adaptive optics systems and the next generation of ground- and space-based telescopes with smaller inner working angles and deeper detection limits will increase the pace of discovery to ultimately map the demographics, composition, evolution, and origin of planets spanning a broad range of masses and ages. | astro-ph.EP | astro-ph |
Draft version May 11, 2016
Preprint typeset using LATEX style emulateapj v. 5/2/11
IMAGING EXTRASOLAR GIANT PLANETS
Brendan P. Bowler1,2
Draft version May 11, 2016
ABSTRACT
High-contrast adaptive optics imaging is a powerful technique to probe the architectures of planetary
systems from the outside-in and survey the atmospheres of self-luminous giant planets. Direct imaging
has rapidly matured over the past decade and especially the last few years with the advent of high-order
adaptive optics systems, dedicated planet-finding instruments with specialized coronagraphs, and
innovative observing and post-processing strategies to suppress speckle noise. This review summarizes
recent progress in high-contrast imaging with particular emphasis on observational results, discoveries
near and below the deuterium-burning limit, and a practical overview of large-scale surveys and
dedicated instruments.
I conclude with a statistical meta-analysis of deep imaging surveys in the
literature. Based on observations of 384 unique and single young (≈5–300 Myr) stars spanning stellar
masses between 0.1–3.0 M⊙, the overall occurrence rate of 5–13 MJup companions at orbital distances
of 30–300 AU is 0.6+0.7
−0.5% assuming hot-start evolutionary models. The most massive giant planets
regularly accessible to direct imaging are about as rare as hot Jupiters are around Sun-like stars.
Dividing this sample into individual stellar mass bins does not reveal any statistically-significant
trend in planet frequency with host mass: giant planets are found around 2.8+3.7
−2.3% of BA stars,
<4.1% of FGK stars, and <3.9% of M dwarfs. Looking forward, extreme adaptive optics systems
and the next generation of ground- and space-based telescopes with smaller inner working angles
and deeper detection limits will increase the pace of discovery to ultimately map the demographics,
composition, evolution, and origin of planets spanning a broad range of masses and ages.
Subject headings: planets and satellites: detection - planets and satellites: gaseous planets
1. INTRODUCTION
Over the past two decades the orbital architecture of
giant planets has expanded from a single order of mag-
nitude in the Solar System (5–30 AU) to over five or-
ders of magnitude among extrasolar planetary systems
(0.01–5000 AU; Figure 1). High-contrast adaptive op-
tics (AO) imaging has played a critical role in this ad-
vancement by probing separations beyond ∼10 AU and
masses &1 MJup. Uncovering planetary-mass objects at
hundreds and thousands of AU has fueled novel theo-
ries of planet formation and migration, inspiring a more
complex framework for the origin of giant planets in
which multiple mechanisms (core accretion, dynamical
scattering, disk instability, and cloud fragmentation) op-
erate on different timescales and orbital separations. In
addition to probing unexplored orbital distances, imag-
ing entails directly capturing photons that originated in
planetary atmospheres, providing unparalleled informa-
tion about the initial conditions, chemical composition,
internal structure, atmospheric dynamics, photospheric
condensates, and physical properties of extrasolar plan-
ets. These three science goals - the architecture, for-
mation, and atmospheres of gas giants - represent the
main motivations to directly image and spectroscopically
characterize extrasolar giant planets.
This pedagogical review summarizes the field of di-
rect imaging in the era leading up to and transition-
ing towards extreme adaptive optics systems, the James
[email protected]
1 McDonald Observatory and the University of Texas at
Austin, Department of Astronomy, 2515 Speedway, Stop C1400,
Austin, TX 78712
2 McDonald Prize Fellow.
Webb Space Telescope, WFIRST, and the thirty meter-
class telescopes. This "classical" period of high-contrast
imaging spanning approximately 2000 to 2015 has set
the stage and baseline expectations for the next gen-
eration of instruments and telescopes that will deliver
ultra-high contrasts and reach unprecedented sensitivi-
ties. In addition to the first images of bona fide extraso-
lar planets, this early phase experienced a number of sur-
prising discoveries including planetary-mass companions
orbiting brown dwarfs; planets on ultra-wide orbits be-
yond 100 AU; enigmatic (and still poorly understood) ob-
jects like the optically-bright companion to Fomalhaut;
and unexpectedly red, methane-free, and dust-rich atmo-
spheres at low surface gravities. Among the most impor-
tant results has been the gradual realization that mas-
sive planets are exceedingly rare on wide orbits; only a
handful of discoveries have been made despite thousands
of hours spent on hundreds of targets spanning over a
dozen surveys. Although dismaying, these null detec-
tions provide valuable information about the efficiency
of planet formation and the resulting demographics at
wide separations. Making use of mostly general, non-
optimized facility instruments and early adaptive optics
systems has also led to creative observing strategies and
post-processing solutions for PSF subtraction.
Distinguishing giant planets from low-mass brown
dwarfs is a well-trodden intellectual exercise (e.g.,
Oppenheimer et al.
2006;
Chabrier et al. 2007; Chabrier et al. 2014).
Except
in the few rare cases where the architectures or abun-
dance patterns of individual systems offer clues about
a specific formation route, untangling the origin of
imaged planetary-mass companions must necessarily
Basri & Brown
2000;
2
75
10
10
1
1
)
p
u
J
M
(
s
s
a
M
1996
2006
2016
0.01 0.1 1 10 100
0.01 0.1 1 10 100
0.01 0.1 1 10 1001000
Separation (AU)
Fig. 1.- Substellar companions discovered via radial velocities (gray circles) and direct imaging (red circles) as of 1996, 2006, and 2016.
Over this twenty year period the number of directly imaged companions below 10 MJup has steadily increased from one (2M1207–3932 b in
2004) to over a dozen. The surprising discovery of planetary companions at extremely wide separations of hundreds to thousands of AU has
expanded the architecture of planetary systems to over five orders of magnitude. Note that the radial velocity planets are minimum masses
(msini) and the directly imaged companion masses are inferred from evolutionary models. RV-detected planets are from exoplanets.org
(Wright et al. 2011; Han et al. 2014) and are supplemented with a compilation of RV-detected brown dwarfs from the literature. Imaged
companions are from Deacon et al. (2014) together with other discoveries from the literature.
be addressed as a population and in a statistical man-
ner. This review is limited in scope to self-luminous
companions detected in thermal emission at near- and
mid-infrared wavelengths (1–5 µm) with masses between
≈1–13 MJup with the understanding that multiple
formation routes can probably produce objects in this
"planetary" mass regime (see Section 5).
Indeed, the
separations regularly probed in high-contrast imaging
surveys- typically tens to hundreds of AU- lie beyond
the regions in protoplanetary disks containing the
highest surface densities of solids where core accretion
operates most efficiently (e.g., Andrews 2015). Direct
imaging has therefore predominantly surveyed the wide
orbital distances where alternative formation and migra-
tion channels like disk instability, cloud fragmentation,
and planet-planet scattering are most likely to apply.
In the future, the most efficient strategy to detect even
smaller super-Earths and terrestrial worlds close to their
host stars will be in reflected light from a dedicated
space-based optical telescope.
recent
2014),
largest
targets,
surveys,
Helling et al.
and statistical
reviews on giant planet
By focusing on the optimal
early dis-
coveries,
results,
this observationally-oriented overview aims to com-
plement
formation
(Chabrier et al. 2007; Helled et al. 2013; Chabrier et al.
2014;
atmospheric models
(Marley et al. 2007a; Helling et al. 2008, Allard et al.
2012; Marley & Robinson 2015), evolutionary mod-
Fortney & Nettelmann
els,
2009),
(Absil & Mawet 2009;
Lagrange 2014; Bailey 2014; Helling & Casewell 2014;
Madhusudhan et al. 2014; Quanz 2015; Crossfield 2015),
and high contrast imaging instruments and speckle
suppression techniques (Guyon et al. 2006; Beuzit et al.
2007; Oppenheimer & Hinkley 2009; Biller et al. 2008;
(Burrows et al.
observational
2001;
results
2010a;
Marois et al.
Mawet et al. 2012b; Davies & Kasper 2012).
2. OPTIMAL TARGETS FOR HIGH-CONTRAST IMAGING
Traub & Oppenheimer
2010;
Planets radiatively cool over time by endlessly releas-
ing the latent heat generated during their formation
and gravitational contraction. Fundamental scaling rela-
tions for the evolution of brown dwarfs and giant plan-
ets can be derived analytically with basic assumptions
of a polytropic equation of state and degenerate electron
gas (Stevenson 1991; Burrows & Liebert 1993). Neglect-
ing the influence of lithium burning, deuterium burning,
and atmospheres, which act as partly opaque wavelength-
dependent boundary conditions, substellar objects with
different masses cool in a similar monotonic fashion over
time:
Lbol ∝ t−5/4M 5/2.
(1)
Here Lbol is the bolometric luminosity, t is the object's
age, and M is its mass.
This steep mass-luminosity relationship means that
luminosity tracks are compressed in the brown dwarf
regime (≈13–75 MJup) and fan out in the planetary
regime with significant consequences for high-contrast
imaging. A small gain in contrast in the brown dwarf
regime results in a large gain in the limiting detectable
mass, whereas the same contrast gain in the planetary
regime has a much smaller influence on limiting mass
(Figure 2).
It is much more difficult, for example, to
improve sensitivity from 10 MJup to 1 MJup than from
80 MJup to 10 MJup. Moreover, sensitivity to low masses
and close separations is highly dependent on a star's
youth and proximity.
In terms of limiting detectable
planet mass, observing younger and closer stars is equiv-
alent to improving speckle suppression or integrating for
3
100
10
1
)
p
u
J
M
(
s
s
a
M
t
e
n
a
P
l
−10 mag
−5 mag
−2 mag
NICI
+2 mag
+5 mag
10
100
5 Gyr
1 Gyr
500 Myr
100 Myr
30 Myr
10 Myr
5 Myr
1
0 p
c
3
0 p
c
1
0
0 p
c
10
Semimajor Axis (AU)
100
10
100
Fig. 2.- The influence of contrast (left), age (middle), and distance (right) on mass sensitivity to planets. The bold curve in each panel
shows the 50% sensitivity contour based on the median NICI contrast from Biller et al. (2013) for a 30 Myr K1 star at 30 pc. The left
panel shows the effect of increasing or decreasing the fiducial contrast curve between –10 magnitudes to +5 magnitudes. Similarly, the
middle and right panels show changes to the fiducial age spanning 5 Myr to 5 Gyr and distances spanning 10 pc to 100 pc. Planet absolute
magnitudes depend steeply on mass and age. As a result, a small gain in contrast in the brown dwarf regime corresponds to a large gain
in limiting mass, but the same contrast gain in the planetary regime translates into a much smaller gain in mass. Mass sensitivity is
particularly sensitive to stellar age, while closer distances mean smaller physical separations can be studied.
longer. Note that in a contrast-limited regime the abso-
lute magnitude of the host star is also important. The
same contrast around low-mass stars and brown dwarfs
corresponds to lower limiting masses compared to higher-
mass stars.
Young stars are therefore attractive targets for two
principal reasons: planets are their most luminous at
early ages, and the relative contrast between young gi-
ant planets and their host stars is lower than at older
ages because stellar luminosities plateau on the main se-
quence while planets and brown dwarfs continue to cool,
creating a luminosity bifurcation. For example, evolu-
tionary models predict the H-band contrast between a
5 MJup planet orbiting a 1 M⊙ star to be ≈25 mag at
5 Gyr but only ≈10 mag at 10 Myr (Baraffe et al. 2003;
Baraffe et al. 2015). At old ages beyond ∼1 Gyr, 1–10
MJup planets are expected to have effective temperatures
between 100–500 K and cool to the late-T and Y spectral
classes with near-infrared absolute magnitudes &18 mag
(Dupuy & Kraus 2013).
Below are overviews of the most common classes of
targets in direct imaging surveys highlighting the scien-
tific context, strengths and drawbacks, and observational
results for each category.
2.1. Young Moving Group Members
In principle, younger stars make better targets for
imaging planets. In practice, the youngest T Tauri stars
reside in star-forming regions beyond 100 pc. At these
distances, the typical angular scales over which high-
contrast imaging can probe planetary masses translate to
wide physical separations beyond ∼20–50 AU (with some
notable exceptions with non-redundant aperture masking
and extreme AO systems). Moreover, these extremely
young ages of ∼1–10 Myr correspond to timescales
when giant planets may still be assembling through core
accretion and therefore might have lower luminosities
than at slightly later epochs (e.g., Marley et al. 2007b;
Molli`ere & Mordasini 2012; Marleau & Cumming 2013).
On the other hand, the closest stars to the Sun probe
the smallest physical scales but their old ages of ∼1–10
Gyr mean that high contrast imaging only reaches brown
dwarf masses.
Young moving groups- coeval, kinematically comov-
ing associations young stars and brown dwarfs in the
solar neighborhood- represent a compromise in age
(≈10–150 Myr) and distance (≈10–100 pc) between
the nearest star forming regions and field stars (Fig-
ure 3; see Zuckerman & Song 2004, Torres et al. 2008,
and Mamajek 2016). One distinct advantage they hold is
that their members span a wide range of masses and can
be used to age date each cluster from lithium depletion
boundaries and isochrone fitting (e.g., Bell et al. 2015;
Herczeg & Hillenbrand 2015). As a result, the ages of
these groups are generally much better constrained than
those of isolated young stars. For these reasons young
moving group members have emerged as the primary tar-
gets for high-contrast imaging planet searches over the
past decade (e.g., Chauvin et al. 2010; Biller et al. 2013;
Brandt et al. 2014c)
Identifying these nearby unbound associations of
young stars is a difficult task. Each moving group's
U V W space velocities cluster closely together with small
velocity dispersions of ≈1–2 km s−1 but individual stars
in the same group can be separated by tens of parsecs
in space and tens of degrees across the sky. U V W kine-
matics can be precisely determined if the proper motion,
radial velocity, and parallax to a star are known.
In-
complete knowledge of one or more of these parameters
(usually the radial velocity and/or distance) means the
U V W kinematics are only partially constrained, mak-
ing it challenging to unambiguously associate stars with
T Tauri Stars
100
5 Myr, 150 pc
Young Moving Groups
30 Myr, 30 pc
0
.
9
4
)
p
u
J
M
(
s
s
a
M
t
e
n
a
P
l
10
1
0
.
0
9
.
1
Field Stars
5 Gyr, 10 pc
0
.
1
9
.
0
1.0
0.8
0.6
0.4
0.2
0.0
D
e
t
e
c
t
i
o
n
P
r
o
b
a
b
i
l
i
t
y
0
.
1
0.1
1.0
100
10
0.1
1.0
100
0.1
Semimajor Axis (AU)
10
1.0
10
100 1000
Fig. 3.- Typical sensitivity maps for high-contrast imaging observations of T Tauri stars (5 Myr at 150 pc), young moving group
members (30 Myr at 30 pc), and field stars (5 Gyr at 10 pc). Young moving group members are "Golidlocks targets"- not too old, not
too distant. Black curves denote 10% and 90% contour levels assuming circular orbits, Cond hot-start evolutionary models (Baraffe et al.
2003), and the median NICI contrast curve from Biller et al. (2013). Gray and orange circles are RV- and directly imaged companions,
respectively (see Figure 1).
r
e
b
m
u
N
500
500
400
400
300
300
200
200
100
100
0
0
YMG Members
Pre−2010
Post−2010
B
A
F
G
K M
Spectral Type
Fig. 4.- The census of members and candidates of young moving
groups. Prior to 2010 the M dwarf members were largely missing
owing to their faintness and lack of parallax measurements from
Hipparcos. Concerted efforts to find low-mass members over the
past few years have filled in this population and generated a wealth
of targets for dedicated direct imaging planet searches.
known groups. Historically, most groups themselves and
new members of these groups were found with the aid of
the Tycho Catalog and Hipparcos, which provided com-
plete space velocities for bright stars together with ancil-
lary information pointing to youth such as infrared ex-
cess from IRAS; X-ray emission from the Einstein or
ROSAT space observatories; strong Hα emission; and/or
Li I λ6708 absorption. As a result, most of the faint low-
mass stars and brown dwarfs have been neglected.
In recent years the population of "missing" low-mass
stars and brown dwarfs in young moving groups has
been increasingly uncovered as a result of large all-sky
dedicated searches (Figure 4; Shkolnik et al. 2009;
2013;
Rodriguez et al.
L´epine & Simon 2009; Schlieder et al. 2010; Kiss et al.
2010; Rodriguez et al. 2011a; Schlieder et al. 2012a;
Schlieder et al. 2012b; Shkolnik et al. 2012; Malo et al.
2013; Moor et al.
2013;
Malo et al. 2014; Gagn´e et al. 2014; Riedel et al. 2014;
Kraus et al. 2014b; Gagn´e et al. 2015b; Gagn´e et al.
2015c; Binks et al. 2015). Parallaxes and radial ve-
locities are generally not available for these otherwise
anonymous objects, but by adopting the U V W kine-
matics of known groups,
it is possible to invert the
problem and predict a distance, radial velocity, and
membership probability. Radial velocities are obser-
vationally cheaper to acquire en masse compared to
parallaxes, so membership confirmation has typically
been accomplished with high-resolution spectroscopy.
The exceptions are for spectroscopic binaries, which
require multiple epochs to measure a systemic velocity,
and rapidly rotating stars with high projected rotational
velocities (vsini), which produce large uncertainties
in radial velocity measurements. The abundance of
low-mass stars in the field means that some old in-
terlopers will inevitably share similar space velocities
with young moving groups. These must be distilled
from bona fide membership lists on a case-by-case basis
(Barenfeld et al. 2013; Wollert et al. 2014; Janson et al.
2014; Mccarthy & Wilhelm 2014; Bowler et al. 2015c).
The current census of directly imaged planets and
companions near the deuterium-burning limit are listed
in Table 1. Many of these host stars are members
of young moving groups. β Pic, 51 Eri, and possi-
bly TYC 9486-927-1 are members of the β Pic mov-
ing group (Zuckerman et al. 2001; Feigelson et al. 2006;
Deacon et al. 2016). HR 8799 and possibly κ And are
thought to be members of Columba (Zuckerman et al.
2011). 2M1207–3932 is in the TW Hydrae Association
(Gizis 2002). GU Psc and 2M0122–2439 are likely mem-
bers of the AB Dor moving group (Malo et al. 2013;
Naud et al. 2014; Bowler et al. 2013). AB Pic, 2M0103–
5515, and 2M0219–3925 are in Tuc-Hor (Song et al. 2003;
Delorme et al. 2013; Gagn´e et al. 2015b), though the
masses of their companions are somewhat uncertain and
may not reside in the planetary regime. In addition, the
space motion of VHS 1256–1257 is well-aligned with the
β Pic or possibly AB Dor moving groups (Gauza et al.
2015; Stone et al. 2016), but the lack of lithium in the
host indicates the system is older and may be a kinematic
interloper.
The number of moving groups in the solar neighbor-
hood is still under debate, but at least five are generally
considered to be well-established: the TW Hydrae As-
sociation, β Pic, Tuc-Hor, Carina, and AB Dor. Other
associations have been proposed and may constitute real
groups which formed together and are useful for age-
dating purposes, but may require more scrutiny to better
understand their size, structure, physical nature, and re-
lationship to other groups. Mamajek (2016) provide a
concise up-to-date census of their status and certitude.
Soon, micro-arcsecond astrometry and parallaxes from
Gaia will dramatically change the landscape of nearby
young moving groups by readily identifying overlooked
groups, missing members, and even massive planets on
moderate orbits.
2.2. T Tauri Stars, Herbig Ae/Be Stars, and
Transition Disks
Despite their greater distances (≈120–150 pc), the ex-
treme youth (≈1–10 Myr) of T Tauri stars and their
massive counterparts, Herbig Ae/Be stars,
in nearby
star-forming regions like Taurus, the Sco-Cen complex,
and ρ Oph have made them attractive targets to search
for planets with direct imaging and probe the ear-
liest stages of planet formation when gas giants are
still assembling (Itoh et al. 2008;
Ireland et al. 2011;
Mawet et al. 2012a; Janson et al. 2013a; Lafreni`ere et al.
2014; Daemgen et al. 2015; Quanz 2015; Hinkley et al.
2015b).
One of the most surprising results from these ef-
forts has been the unexpected discovery of planetary-
mass companions on ultra-wide orbits at several hundred
AU from their host stars (Table 1). These wide com-
panions pose challenges to canonical theories of planet
formation via core accretion and disk instability and
may instead represent the tail end of brown dwarf
companion formation, perhaps as opacity-limited frag-
ments of turbulent, collapsing molecular cloud cores (e.g.,
Low & Lynden-Bell 1976; Silk 1977; Boss 2001; Bate
2009).
Many (and perhaps most) of these young planetary-
mass companions harbor accreting circum-planetary
disks, which provide valuable information about mass
accretion rates, circum-planetary disk structure, forma-
tion route, and the moon-forming capabilities of young
planets. Accretion luminosity is partially radiated in
line emission, making Hα a potentially valuable tracer to
find and characterize protoplanets (Sallum et al. 2015a).
For example, Zhou et al. (2014) find that up to 50% of
the accretion luminosity in the ≈15 MJup companion
GSC 6214-210 B is emitted at Hα. Searching for these
nascent protoplanets has become a leading motivation to
achieve AO correction in the optical and is actively being
carried out with MagAO (Close et al. 2014). Deep sub-
5
mm observations with ALMA have opened the possibility
of measuring the masses of these subdisks (Bowler et al.
2015a) and possibly even indirect identification via gas
kinematics (Perez et al. 2015). Larger disks may be able
to be spatially resolved and a dynamical mass for the
planet may be measured from Keplerian motion.
The relationship between protoplanetary disks and
young planets is also being explored in detail at these
extremely young ages. In particular, transition disks-
young stars whose spectral energy distributions indicate
they host disks with large optically thin cavities generally
depleted of dust (e.g., see reviews by Williams & Cieza
2011, Espaillat et al. 2014, Andrews 2015, and Owen
2016)- have been used as signposts to search for em-
bedded protoplanets. This approach has been quite fruit-
ful, resulting in the discovery of companions within these
gaps spanning the stellar (CoKu Tau 4: Ireland & Kraus
2008; HD 142527:
Biller et al. 2012, Close et al.
2014, Rodigas et al. 2014, Lacour et al. 2015), brown
dwarf (HD 169142: Biller et al. 2014, Reggiani et al.
2014), and planetary (LkCa 15: Kraus & Ireland 2012,
Ireland & Kraus 2014, Sallum et al. 2015a; HD 100546:
Quanz et al. 2013, Currie et al. 2014d, Quanz et al.
2015; Currie et al. 2015, Garufi et al. 2016) mass regimes
using a variety of techniques.
However, environmental factors can severely compli-
cate the interpretation of these detections. Extinction
and reddening, accretion onto and from circum-planetary
disks, extended emission from accretion streams, and
circumstellar disk sub-structures seen in scattered light
can result in false alarms, degenerate interpretations,
and large uncertainties in the mass estimates of ac-
tual companions. T Cha offers a cautionary example;
Hu´elamo et al. (2011) discovered a candidate substel-
lar companion a mere 62 mas from the transition-disk
host star with aperture masking interferometry, but ad-
ditional observations did not show orbital motion as ex-
pected for a real companion. Additional modeling in-
dicates that the signal may instead be a result of scat-
tering by grains in the outer disk or possibly even noise
in the data (Olofsson et al. 2013; Sallum et al. 2015b;
Cheetham et al. 2015). This highlights an additional
complication with aperture masking: because model fits
to closure phases usually consist of binary models with
two or more point sources, it can difficult to discern ac-
tual planets from other false positives. In these situations
the astrometric detection of orbital motion is essential to
confirm young protoplanets embedded in disks.
Other notable examples of ambiguous candidate pro-
toplanets at wider separations include FW Tau b, an
accreting low-mass companion to the Taurus binary
FW Tau AB orbiting at a projected separation of 330 AU
(White & Ghez 2001; Kraus et al. 2014a), and TMR-
1C, a faint, heavily extincted protoplanet candidate lo-
cated ≈1400 AU from the Taurus protostellar binary
host TMR-1AB showing large-amplitude photometric
variability and circumstantial evidence of a dynami-
cal ejection (Terebey et al. 1998; Terebey et al. 2000;
Riaz & Mart´ın 2011; Riaz et al. 2013). Follow-up ob-
servations of both companions suggest they may in-
stead be low-mass stars or brown dwarfs with edge-
on disks (Petr-Gotzens et al. 2010; Bowler et al. 2014;
Kraus et al. 2015; Caceres et al. 2015), underscoring a
few of the difficulties that arise when interpreting candi-
6
date protoplanets at extremely young ages.
Altogether
the statistics of planets orbiting the
youngest T Tauri stars from direct imaging are still fairly
poorly constrained. Quantifying this occurrence rate is
important because it can be compared with the same val-
ues at older ages to determine the evolution of this popu-
lation. Planet-planet scattering, for example, implies an
increase in the frequency of planets on ultra-wide orbits
over time. Ireland et al. (2011) found the frequency of 6–
20 MJup companions from ≈200–500 AU to be ∼4+5
−1%
in Upper Scorpius. Combining these results with those
from Kraus et al. (2008) and their own shallow imaging
survey, Lafreni`ere et al. (2014) find that the frequency
of 5–40 MJup companions between 50–250 AU is <1.8%
and between 250–1000 AU is 4.0+3.0
−1.2% assuming hot-start
evolutionary models. In future surveys it will be just as
important to report nondetections together with new dis-
coveries so this frequency can be measured with greater
precision.
2.3. Brown Dwarfs
Young brown dwarfs
(≈13–75 MJup) have
low
circum-substellar disk masses
(Mohanty et al. 2013;
Andrews et al. 2013) and are not expected to host giant
planets as frequently as stars. Nevertheless, their low lu-
minosities make them especially advantageous for high-
contrast imaging because lower masses can be probed
with contrast-limited observations.
(Kraus et al.
2005;
Ahmic et al.
Several deep imaging surveys with ground-based
AO or HST have included brown dwarfs in their
samples
2007;
Stumpf et al. 2010; Biller et al. 2011; Todorov et al.
2014; Garcia et al. 2015). A handful of companions
in the 5–15 MJup range have been discovered with
direct imaging: 2M1207–3932 b (Chauvin et al. 2004;
Chauvin et al. 2005a), 2M0441+2301 b (Todorov et al.
2010), and possibly both FU Tau B (Luhman et al.
2009b)
2015)
depending on their ages. A few other low-mass com-
panions (and in some cases the primaries themselves)
to late-T and early-Y field brown dwarfs may also
reside in the planetary regime depending on the
system ages: CFBDSIR J1458+1013 B (Liu et al.
2011), WISE J1217+1626 B (Liu et al. 2012), and
WISE J0146+4234 B (Dupuy et al. 2015a).
and VHS 1256–1257
(Gauza et al.
The low mass ratios of these systems (q≈0.2–0.5) bear
a closer resemblance to binary stars than canonical plan-
etary systems (q.0.001), and the formation route of
these very low-mass binaries is probably quite different
than around stars (Lodato et al. 2005). High-order mul-
tiple systems with low total masses like 2M0441+2301
AabBab and VHS 1256–1257 ABb suggest that cloud
fragmentation can form objects in the planetary-mass
domain (Chauvin et al. 2005a; Todorov et al. 2010;
Bowler & Hillenbrand 2015; Stone et al. 2016). Con-
tinued astrometric monitoring of ultracool binaries will
eventually yield orbital elements and dynamical masses
for these intriguing systems to test formation mecha-
nisms (Dupuy & Liu 2011) and giant planet evolutionary
models.
2.4. Binary Stars
Close stellar binaries (≈0.1–5′′) are generally avoided
in direct imaging surveys. Multiple similarly-bright
stars can confuse wavefront sensors, which are opti-
mized for single point sources, and deep coronagraphic
imaging generally saturates nearby stellar companions.
Physically, binaries carve out large dynamically-unstable
regions that are inhospitable to planets and there is
strong evidence that they inhibit planet formation by
rapidly clearing or truncating protoplanetary disks (e.g.,
Cieza et al. 2009; Duchene 2010; Kraus et al. 2012).
Nevertheless, many planets have been found in binary
systems in both S-type orbital configurations (a planet
orbiting a single star; e.g., Ngo et al. 2015) and P-type
orbits (circumbinary planets; see Winn & Fabrycky 2015
for a recent summary). Binaries are common products
of star formation, so understanding how stellar multiplic-
ity influences the initial conditions (protoplanetary disk
mass and structure), secular evolution (Kozai-Lidov in-
teractions) and end products (dynamically relaxed plan-
etary systems) of planet formation has important con-
sequences for the galactic census of exoplanets (e.g.,
Wang et al. 2014; Kraus et al. 2016).
Several planetary-mass companions have been im-
aged around binary stars on wide circumbinary orbits:
ROXs 42B b (Kraus et al. 2014a; Currie et al. 2014b),
Ross 458 c (Goldman et al. 2010; Scholz 2010), SR 12 C
(Kuzuhara et al. 2011), HD 106906 b (Bailey et al. 2014;
Lagrange et al. 2016), and VHS 1256–1257 (Gauza et al.
2015; Stone et al. 2016). 2M0103–5515 b (Delorme et al.
2013) and FW Tau b (Kraus et al. 2014a) orbit close,
near-equal mass stellar binaries, but the masses of the
wide tertiaries are highly uncertain (Bowler et al. 2014).
On the other hand, few imaged planets orbiting sin-
gle stars also have wide stellar companions. 51 Eri Ab
is orbited by the pair of M dwarf binaries GJ 3305 AB
(Feigelson et al. 2006; Kasper et al. 2007; Montet et al.
2015) at ∼2000 AU. Fomalhaut has two extremely dis-
tant stellar companions at ≈57 kAU and ≈158 kAU
(Mamajek et al. 2013). 2M0441+2301 Bb orbits a low-
mass brown dwarf and is part of a hierarchical quadruple
system with a distant star-brown dwarf pair at a pro-
jected separation of 1800 AU (Todorov et al. 2010).
There has been little work comparing the occurrence
rate of imaged planets in binaries and single stars. How-
ever, several surveys and post-processing techniques are
now expressly focusing on binary systems and should
clarify the statistical properties of planets in these dy-
namically complicated arrangements (Thalmann et al.
2014; Rodigas et al. 2015; Thomas et al. 2015).
2.5. Debris Disks: Signposts of Planet Formation?
Debris disks are extrasolar analogs of the asteroid
and Kuiper belts in the Solar System. They are the
continually-replenished outcomes of cascading planetesi-
mal collisions that result in large quantities of transient
dust heated by the host star. Observationally, debris
disks are identified from unresolved infrared or sub-mm
excesses over stellar photospheric emission. Deep ob-
servations spanning the optical, IR, and sub-mm can
spatially resolved the largest and most luminous disks
in scattered light and thermal emission to investigate
disk morphology and grain properties. Topical reviews
by Zuckerman (2001), Moro-Martin et al. (2008), Wyatt
(2008), Krivov (2010), and Matthews et al. (2014a) high-
light recent theoretical and observational progress on the
formation, modeling, and evolution of debris disks.
Debris disks are intimately linked to planets, which
can stir planetesimals, sculpt disk features, produce off-
sets between disks and their host stars, and carve gaps
to form belts with spectral energy distributions showing
multiple temperature components. The presence of de-
bris disks, and especially those with features indicative
of a massive perturber, may therefore act as signposts
for planets. The four directly imaged planetary systems
β Pic, HR 8799, 51 Eri, and HD 95086 all possess debris
disks, the latter three having multiple belts interior and
exterior to the imaged planet(s). This remarkably con-
sistent configuration is analogous to the Solar System's
architecture in which gas giants are flanked by (very low-
level) zodiacal emission.
Anecdotal signs point to a possible correlation be-
tween disks and imaged planets but this relationship has
not yet been statistically validated. There have been
hints of a correlation between debris disks and low-mass
planets detected via radial velocities (Wyatt et al. 2012;
Marshall et al. 2014), but these were not confirmed in
a recent analysis by Moro-Martin et al. (2015). Indeed,
many stars hosting multi-component debris disks have
now been targeted with high-contrast imaging and do not
appear to harbor massive planets (Rameau et al. 2013a;
Wahhaj et al. 2013b; Janson et al. 2013c; Meshkat et al.
2015a). Given the high incidence of debris disks around
main-sequence stars (&16–20%: Trilling et al. 2008;
Eiroa et al. 2013), with even higher rates at younger ages
(Rieke et al. 2005; Meyer et al. 2008), any correlation of
imaged giant planets and debris disks will be difficult to
discern because the overall occurrence rate of massive
planets on wide orbits is extremely low (.1%; see Sec-
tion 4.5). Perhaps more intriguing would be a subset
of this sample with additional contextual clues, for ex-
ample the probability of an imaged planet given a two-
component debris disk compared to a diskless control
sample.
The Fabulous Four- Vega, β Pic, Fomalhaut, and
ǫ Eridani- host the brightest debris disks discovered
by IRAS (Aumann et al. 1984; Aumann 1985) and have
probably been targeted more than any other stars with
high-contrast imaging over the past 15 years, except per-
haps for HR 8799. Despite having similarly large and lu-
minous disks, their planetary systems are quite different
and demonstrate a wide diversity of evolutionary out-
comes.
Fomalhaut's disk possesses a sharply truncated, off-
set, and eccentric ring about 140 AU in radius sug-
(Dent et al. 2000;
gesting sculpting from a planet
Boley et al. 2012; Kalas et al. 2005).
A comoving
optical source ("Fomalhaut b") was discovered inte-
rior to the ring by Kalas et al. (2008a) and appears
to be orbiting on a highly inclined and eccentric or-
bit not coincident with the ring structure (Kalas et al.
2013). The nature of this intriguing companion re-
mains puzzling;
it may be a low-mass planet with a
large circum-planetary disk, a swarm of colliding irregu-
lar satellites, or perhaps a recent collision of protoplan-
ets (e.g., Kalas et al. 2008b; Kennedy & Wyatt 2011;
Kenyon et al. 2014). Massive planets have been ruled out
from deep imaging down to about 20 AU (Kalas et al.
2008a; Kenworthy et al. 2009; Marengo et al. 2009;
7
Absil et al. 2011; Janson et al. 2012; Nielsen et al. 2013;
Kenworthy et al. 2013; Currie et al. 2012b; Currie et al.
2013; Janson et al. 2015).
Vega's nearly face-on debris disk is similar to Foma-
lhaut's in terms of its two-component structure com-
prised of warm and cold dust belts and wide gaps
with orbital ratios &10, possibly indicating the pres-
ence of multiple low-mass planets (e.g., Wilner et al.
2002; Su et al. 2005; Su et al. 2013). Deep imaging of
Vega over the past 15 years has thus far failed to iden-
tify planets with detection limits down to a few Jupiter
masses (Metchev et al. 2003; Macintosh et al. 2003;
Itoh et al. 2006; Marois et al. 2006; Hinz et al. 2006;
Hinkley et al. 2007; Heinze et al. 2008; Janson et al.
2011a; Mennesson et al. 2011; Janson et al. 2015).
β Pic hosts an extraordinarily large, nearly edge-on
disk spanning almost 2000 AU in radius (Smith & Terrile
1984; Larwood & Kalas 2001). Its proximity, brightness,
and spatial extent make it one of the best-studied debris
disks, showing signs of multiple belts (e.g., Wahhaj et al.
2003), asymmetries (e.g., Kalas & Jewitt 1995), molec-
ular gas clumps (Dent et al. 2014), and an inner warp
(Heap et al. 2000) predicted to be caused by a close-
in inclined planet (Mouillet et al. 1997). Lagrange et al.
(2009a) uncovered a possible massive planet at ∼9 AU
and, despite immediate follow-up (Fitzgerald et al. 2009;
Lagrange et al. 2009b), it was not until it reemerged on
the other side of the star that β Pic b was unambiguously
confirmed (Lagrange et al. 2010).
ǫ Eridani
is another particularly fascinating exam-
ple of a nearby, relatively young K2 star hosting a
bright debris disk with spatially resolved ring struc-
ture and a warm inner component (Greaves et al. 1998;
Greaves et al. 2005; Backman et al. 2009). At 3.2 pc this
star harbors the closest debris disk to the Sun, has a
Jovian-mass planet detected by radial velocity and as-
trometric variations (Hatzes et al. 2000; Benedict et al.
2006), and possesses a long-term RV trend pointing to
an additional long-period giant planet. Because of its
favorable age and proximity, ǫ Eridani has been exhaus-
tively imaged with adaptive optics on the largest ground-
based telescopes in an effort to recover the known planets
and search for others (Luhman & Jayawardhana 2002;
Macintosh et al. 2003; Itoh et al. 2006; Marengo et al.
2006;
Lafreni`ere et al. 2007b;
Biller et al. 2007; Janson et al. 2008; Heinze et al. 2008;
Marengo et al. 2009; Heinze et al. 2010b; Wahhaj et al.
2013b; Janson et al. 2015). Together with long-baseline
radial velocity monitoring, these deep observations have
ruled out planets >3 MJup anywhere in this system.
Janson et al. 2007;
2.6. Field Stars and Radial Velocity Trends
At the old ages of field stars (∼1–10 Gyr), giant plan-
ets have cooled to late spectral types and low luminosi-
ties where high-contrast imaging does not regularly reach
the planetary-mass regime. Nevertheless, several surveys
have focused on this population because their proximity
means very close separations can be probed and their old
ages provide information about potential dynamical evo-
lution of substellar companions and giant planets over
time (McCarthy & Zuckerman 2004; Carson et al. 2005;
Carson et al. 2006; Heinze et al. 2010a; Heinze et al.
2010b; Tanner et al. 2010; Leconte et al. 2010).
Of particular interest are stars showing low-amplitude,
8
long-baseline radial velocity changes (Doppler "trends").
These accelerations are regularly revealed in planet
searches and point to the existence of unseen stars,
brown dwarfs, or giant planets on wide orbits. High-
contrast imaging is a useful tool to diagnose the na-
ture of these companions and,
in the case of non-
detections, rule out massive objects at wide projected
separations (Kasper et al. 2007; Geissler et al. 2007;
Luhman & Jayawardhana 2002; Chauvin et al. 2006;
Janson et al. 2009; Jenkins et al. 2010; Kenworthy et al.
2009; Rodigas et al. 2011).
When a companion is detected, its minimum mass can
be inferred from the host star's acceleration ( v), the dis-
tance to the system (d), and the angular separation of
the companion (ρ) following Torres (1999) and Liu et al.
(2002):
Mcomp > 1.388 × 10−5(cid:18) d
pc
ρ
′′(cid:19)2
M⊙. (2)
v
(cid:12)(cid:12)(cid:12)
m s−1 yr−1(cid:12)(cid:12)(cid:12)
The coefficient is 0.0145 when expressed in MJup. This
equation assumes an instantaneous radial velocity slope,
but longer baseline coverage or a change in the accelera-
tion ("jerk") can provide better constraints on a com-
panion's mass and period (Rodigas et al. 2016).
If a
significant fraction of the orbit is measured with both
astrometry and radial velocities, simultaneous model-
ing of both data sets can yield a robust dynamical
mass measurement. Perhaps the best example of this is
from Crepp et al. (2012a), who measured the mass and
three-dimensional orbit of the brown dwarf companion
HR 7672 B, initially discovered by Liu et al. (2002) based
on an acceleration from the host star.
Many stellar and white dwarf companions have been
discovered in this fashion but only a few substellar com-
panions have been found (Table 2). Figure 5 shows the
population of known companions inducing shallow RV
trends on their host stars and which have also been re-
covered with high resolution (and often high-contrast)
imaging. Most of these are M dwarfs with masses be-
tween ∼0.1–0.5 M⊙ at separations of ∼10–100 AU. This
is primarily due to the two competing methods at play:
at these old ages, direct imaging is insensitive to low
masses and close separations, while small accelerations
induced from wide-separation and low-mass companions
are difficult to measure even for long-baseline, precision
radial velocity planet searches. The TRENDS program
(e.g., Crepp et al. 2012a; Crepp et al. 2014) is the largest
survey to combine these two methods and demonstrates
the importance of both detections and non-detections to
infer the population of planets on moderate orbits out to
∼20 AU (Montet et al. 2014).
Dynamical masses of planets may eventually be mea-
sured by combing radial velocity monitoring of the host
star and direct imaging, effectively treating the system
like a spatially resolved single-lined spectroscopic bi-
nary. Stellar jitter is a limiting factor at very young
ages and at older ages the low luminosities of planets
generally preclude imaging. The intermediate ages of
moving group members may provide an adequate solu-
tion, and at least one ambitious survey by Lagrange et al.
(2013) is currently underway to search for planets and
long-term radial velocity trends for this population. An-
1.00
)
n
u
S
M
(
s
s
a
M
n
o
n
a
p
m
o
C
i
0.10
30 m s−1 yr−1
HR 7672 B
HD 4747 B
HIP 71898 B
HD 19467 B
Substellar Limit
10 m s−1 yr−1
0.01
10
3 m s−1 yr−1
1 m s−1 yr−1
Larger
Acceleration
Easier to
Image
Harder to
Image
Smaller
Acceleration
White Dwarf
Low−Mass Star
Brown Dwarf
Planet Host
Separation (AU)
100
Fig. 5.- Imaged companions inducing shallow radial veloc-
ity trends on their host stars. Blue, orange, and red circles are
white dwarf companions, low-mass stellar companions, and sub-
stellar companions, respectively. Concentric circles indicate the
host star has a planetary system. Only three brown dwarf com-
panions inducing shallow trends have been found: HR 7672 B
(Liu et al. 2002), HD 19467 B (Crepp et al. 2014), and HD 4747 B
(Crepp et al. 2016). Gray dashed lines show constant accelera-
tions assuming circular orbits; the maximum host star acceleration
is proportional to the companion mass and inversely proportional
to the square of the projected physical separation, so a 1 MJup
planet at 10 AU will produce the same maximum acceleration as a
100 MJup low-mass star at 100 AU (namely 0.7 m s−1 yr−1). See
Table 2 for details on these systems.
other solution is to image planets in reflected light at
optical wavelengths, which requires a space-based tele-
scope and coronagraph like WFIRST (Traub et al. 2014;
Spergel et al. 2015; Brown 2015; Greco & Burrows 2015;
Robinson et al. 2016). Similarly, astrometric accelera-
tions can be used to identify and measure the masses
of substellar companions when combined with high-
contrast imaging. This will be particularly relevant in
the near-future with Gaia as thousands of planets are ex-
pected to be found from the orbital reflex motion of their
host stars (Sozzetti et al. 2013; Perryman et al. 2014).
Young stars in the field not necessarily associated
with coherent moving groups have also been popu-
lar targets for high-contrast imaging planet searches.
The advantage of this population is that they are nu-
merous and often reside at closer distances than ac-
tual members of young moving groups, but their ages
and metallicities are generally highly uncertain so sub-
stellar companions uncovered with deep imaging can
have a wide range of possible masses (e.g., Mawet et al.
2015). GJ 504 and κ And are recent examples of
young field stars with faint companions that were initially
thought to have planetary masses (Kuzuhara et al. 2013;
Carson et al. 2013) but which follow-up studies showed
are probably older (Hinkley et al. 2013; Bonnefoy et al.
2014a; Fuhrmann & Chini 2015; Jones et al. 2016), im-
plying the companions are likely more massive and re-
side in the brown dwarf regime. Sirius is another ex-
ample of a young massive field star extensively tar-
geted with high-contrast imaging (Kuchner & Brown
2000; Bonnet-Bidaud & Pantin 2008; Skemer & Close
2011; Thalmann et al. 2011; Vigan et al. 2015). This
system is particularly noteworthy for possible periodic
astrometric perturbations to the orbit of its white dwarf
companion Sirius B that may be caused by an still-hidden
giant planet or brown dwarf (Benest & Duvent 1995).
3. THE MASSES OF IMAGED PLANETS
The masses of directly imaged planets are generally
highly uncertain, heavily model-dependent, and diffi-
cult to independently measure. Yet mass is fundamen-
tally important to test models of giant planet formation
and empirically calibrate substellar evolutionary models.
This Section describes how observables like bolometric
luminosity, color, and absolute magnitude coupled with
evolutionary models and semi-empirical quantities like
age are used to infer the masses of planets. Although no
imaged planet has yet had its mass directly measured,
there are several promising routes to achieve this which
will eventually enable rigorous tests of giant planet cool-
ing models.
3.1. Inferring Masses
Like white dwarfs and brown dwarfs, giant planets
cool over time so evolutionary models along with two
physical parameters- luminosity, age, effective temper-
ature, or radius- are needed to infer a planet's mass.
Among these, luminosity and age are usually better con-
strained and less reliant on atmospheric models than ef-
fective temperature and radius, which can substantially
vary with assumptions about cloud properties, chemi-
cal composition, and sources of opacity. Below are sum-
maries of the major assumptions (in roughly descending
order) involved in the inference of planet masses using
atmospheric and evolutionary models along with notable
advantages, drawbacks, and limitations of various tech-
niques.
• Initial conditions and formation pathway.
The most important assumption is the amount of
initial energy and entropy a planet begins with fol-
lowing its formation. This defines its evolutionary
pathway, which is embodied in three broad classes
informed by formation mechanisms.
Hot-start models begin with arbitrarily large
radii and oversimplified,
idealized initial condi-
tions that generally ignore the effects of accre-
tion and mass assembly. As such, they rep-
resent the most luminous outcome and corre-
spond to the most optimistically (albeit unrealis-
tically) low mass estimates.
Ironically, hot-start
grids are nearly unanimously adopted for estimat-
ing the masses of young brown dwarfs and giant
planets even though the early evolution in these
models is the least reliable. The most widely
used hot-start models for imaged planets are the
Cond and Dusty grids from Baraffe et al. (2003)
and Chabrier (2001), Burrows et al. (1997), and
Saumon & Marley (2008).
Cold-start models were made prominent by
Marley et al. (2007b) and Fortney et al. (2008) in
the context of direct imaging as an attempt to em-
ulate a more realistic formation scenario for giant
planets through core accretion. In this model, ac-
cretion shocks radiate the gravitational potential
energy of infalling gas as a giant planet grows.
After formation, these planets begin cooling with
9
much lower luminosities and initial entropies com-
pared to the hot-start scenario, taking between
∼108 and ∼109 years to converge with hot-start
cooling models depending on the planet mass. The
observational implications of this are severe: plan-
ets formed from core accretion may be orders of
magnitude less luminous than those produced from
cloud fragmentation or disk instability. While this
may offer a diagnostic for the formation route if
the mass of a planet is independently measured,
it also introduces considerable uncertainty in the
more typical case when only an age and luminos-
ity are known. For example, 51 Eri b may be as
low as 2 MJup or as high as 12 MJup depending
on which cooling model (hot or cold) is assumed
(Macintosh et al. 2015).
This picture is made even more complicated by
large uncertainties in the details of cold-start mod-
els. The treatment of accretion shocks, circumplan-
etary disks, core mass, and even deuterium burning
for the most massive planets can dramatically influ-
ence the initial entropy and luminosity evolution of
planets (Mordasini et al. 2012; Bodenheimer et al.
2013; Mordasini 2013; Owen & Menou 2016).
This motivated a class of warm-start models
with intermediate initial entropies that prob-
ably better reflect dissipative accretion shocks
that occur in nature (Spiegel & Burrows 2012;
Marleau & Cumming 2013). Unfortunately, the
relevant details of giant planet assembly are poorly
constrained by observations. There is also likely
to be intrinsic scatter in the initial conditions for
a given planet which may result in large degenera-
cies in the planet mass, core mass, and accretion
history for young gas giants with the same age and
luminosity. It is quite possible, for instance, that
the HR 8799 c, d, and e planets which all share
the same age and nearly the same luminosity could
have very different masses.
• Stellar age. After bolometric luminosity, which
is generally uncomplicated to estimate for imaged
planets, the age of the host star is the most sen-
sitive parameter on which the mass of an im-
aged companion depends.
It is also one of the
most difficult quantities to accurately determine
and usually relies on stellar evolutionary models
or empirical calibrations. Recent reviews on this
topic include Soderblom (2010), Jeffries (2014),
and Soderblom et al. (2014). Figure 6 shows the
current census of imaged companions near and be-
low the deuterium-burning limit with both age and
luminosity measurements (from Table 1). Apart
from uncertainties in the formation history of these
objects, age uncertainties dominate the error bud-
get for inferring masses.
Clusters of coeval stars spanning a wide range of
masses provide some of the best age constraints
but are still dominated by systematic errors. Sev-
eral star-forming regions and young moving groups
in particular have been systematically adjusted to
older ages over the past few years, which has prop-
agated to the ages and masses of planets in those
10
/
)
n
u
S
L
L
(
g
o
−5
−6
−7
−2
−3
−4
SR12 C
ROXs42B b
HD 106906 b
βPic b
l
2M1207−39 b HR 8799 cde
HD 95086 b
GU Psc b
HR 8799 b
Ross 458 c
51 Eri b
GJ 758 B
GJ 504 b
CFBDS1458+10 B
WISE1217+16 B
WISE0146+42 B
6
7
8
log(Age) (yr)
9
10
Fig. 6.- The current census of companions in the brown dwarf (green) and planetary (blue) mass regimes that have both age and
bolometric luminosity measurements from the compilation in Table 1. Many companions lie near the deuterium-burning limit while only a
handful of objects are unambiguously in the planetary-mass regime. Hot-start evolutionary models are from Burrows et al. (1997); orange,
green, and blue tracks denote masses >80 MJup, 14–80 MJup, and <14 MJup.
associations (Pecaut et al. 2012; Binks & Jeffries
2014; Kraus et al. 2014b; Bell et al. 2015). The
implied hot-start mass for β Pic b, for example, in-
creases by several Jupiter masses (corresponding to
several tens of percent) assuming the planet's age
is ≈23 Myr instead of ≈12 Myr (Mamajek & Bell
2014; although see the next bullet point).
For young field stars, distant stellar companions
can help age-date the entire system.
For ex-
ample, the age of Fomalhaut was recently re-
vised to ∼400 Myr from ∼200 Myr in part due
to constraints from its wide M dwarf compan-
ions (Mamajek 2012; Mamajek et al. 2013). Ul-
timately, if the age of a host star is unknown, the
significance and interpretation of a faint companion
is limited if basic physical properties like its mass
are poorly constrained.
• Epoch of planet formation. Planets take time
to form so they are not exactly coeval with their
host stars. Their ages may span the stellar age
to the stellar age minus ∼10 Myr depending on
the timescale for giant planets to assemble. Plan-
ets formed via cloud fragmentation or disk insta-
bility might be nearly coeval with their host star,
but those formed by core accretion are expected to
build mass over several Myr. While this difference
is negligible at intermediate and old ages beyond a
few tens of Myrs, it can have a large impact on the
inferred masses of the youngest planets (.20 Myr).
For example, if the age of the young planetary-mass
companion 2M1207–3932 b is assumed to be coeval
with the TW Hyrdae Association (τ = 10 ± 3 Myr)
then its hot-start mass is ≈5 MJup. On the other
hand, if its formation was delayed by 8 Myr (τ =
2 Myr) then its mass is only ≈2.5 MJup.
• Atmospheric models. Atmospheric models can
influence the inferred masses of imaged exoplan-
ets in several ways. They act as surface bound-
ary conditions for evolutionary models and regu-
late radiative cooling through molecular and con-
tinuum opacity sources. This in turn impacts the
luminosity evolution of giant planets, albeit min-
imally because of the weak dependence on mean
opacity (L(t) ∝ κ0.35; Burrows & Liebert 1993;
Burrows et al. 2001). Even the unrealistic cases
of permanently dusty and perpetually condensate-
free photospheres do not dramatically affect the lu-
minosity evolution of cooling models or mass de-
terminations using age and bolometric luminosity
(Baraffe et al. 2002; Saumon & Marley 2008), al-
though more realistic ("hybrid") models account-
ing for the evolution and dissipation of clouds
at the L/T transition can influence the shape
of cooling curves in slight but significant ways
(Saumon & Marley 2008; Dupuy et al. 2015b).
On the other hand, mass determinations in color-
magnitude space are highly sensitive to atmo-
spheric models and can result in changes of several
tens of percent depending on the specific treatment
of atmospheric condensates. Dust reddens spec-
−10
−5
0
5
10
15
20
)
g
a
m
(
H
M
Lum Class I
IV
V
T
Y
II
III
M Dwarfs
L
25
−2
−1
0
J−K (mag)
1
2
3
Fig. 7.- The modern color-magnitude diagram spans nearly 35
magnitudes in the near-infrared and 5 magnitudes in J–K color.
The directly imaged planets (bold circles) extend the L dwarf se-
quence to redder colors and fainter absolute magnitudes owing to
a delayed transition from cloudy atmospheres to condensate-free
T dwarfs at low surface gravities. The details of this transition
for giant planets remains elusive. OBAFGK stars (gray) are from
the extended Hipparcos compilation XHIP (Anderson & Francis
2012); M dwarfs (orange) are from Winters et al. (2015);
late-
M dwarfs (orange), L dwarfs (green), and T dwarfs (light blue)
are from Dupuy & Liu (2012); and Y dwarfs (blue) are com-
piled largely from Dupuy et al. (2014), Tinney et al. (2014), and
Beichman et al. (2014) by T. Dupuy (2016, private communica-
tion). Directly imaged planets or planet candidates (bold circles)
represent all companions from Table 1 with near-infrared photom-
etry and parallactic distances.
tra and can modify the near-infrared colors and
absolute magnitudes of ultracool objects by sev-
eral magnitudes. This introduces another source
of uncertainty if the spectral shape is poorly con-
strained, though the difference between dusty and
cloud-free models is smaller at longer wavelengths
and higher temperatures.
the most
One of
important and unexpected
empirical results to emerge from direct imag-
ing has been the realization that young brown
dwarfs and massive planets retain photospheric
clouds even at low effective temperatures where
older, high-gravity brown dwarfs have already
transitioned to T dwarfs (Metchev & Hillenbrand
2006; Chauvin et al. 2004; Marois et al. 2008;
Patience et al.
2010a;
Faherty et al. 2012; Bowler et al. 2013; Liu et al.
2013; Filippazzo et al. 2015). This is demonstrated
in Figure 7, which shows the location of imaged
Bowler et al.
2010;
11
companions near and below the deuterium-
burning limit on the near-infrared color-color
diagram. At young ages, warm giant planets are
significantly redder than the field population of
brown dwarfs, and several of the most extreme
examples have anomalously low absolute mag-
nitudes. For old brown dwarfs, this evolution
from dusty, CO-bearing L dwarfs to cloud-free,
methane-dominated T dwarfs takes place over a
narrow temperature range (∼1200–1400 K) but
occurs at a lower (albeit still poorly constrained)
temperature for young gas giants. The lack of
methane is likely caused by disequilibrium carbon
chemistry at low surface gravities as a result
of vigorous vertical mixing (e.g., Barman et al.
2011a; Zahnle & Marley 2014;
Ingraham et al.
2014; Skemer et al. 2014a), while the preservation
of photospheric condensates can be explained by
a dependency of cloud base pressure and particle
size on surface gravity (Marley et al. 2012). Un-
fortunately, the dearth of known planets between
∼L5–T5 is the main limitation to understanding
this transition in detail (Figure 8).
In principle, the mass of a planet can also be in-
ferred by fitting synthetic spectra to the planet's
observed spectrum or multi-band photometry. The
mass can then be obtained from best-fitting model
as follows:
Mp (MJup) = 12.76 × 10log(g)−4.5 dex(cid:18) R
RJup(cid:19)2
.
(3)
Here log(g) is the surface gravity (in cm s−2)
and R is the planet's radius. The radius can
either be taken from evolutionary models or al-
ternatively from the multiplicative factor that
scales the emergent model spectrum to the ob-
served flux-calibrated spectrum (or photometry) of
the planet. This scale factor corresponds to the
planet's radius over its distance, squared (R2/d2;
see Cushing et al. 2008 for details).
Clearly the inferred mass is very sensitive to both
the surface gravity and the radius.
In practice,
gravity is usually poorly constrained for model fits
to brown dwarf and giant planet spectra because its
influence on the emergent spectrum is more sub-
tle (e.g., Cushing et al. 2008; Bowler et al. 2011;
Barman et al. 2011a; Macintosh et al. 2015).
In
addition, the scale factor strongly depends on the
model effective temperature (∝ T −4
eff ), which is typ-
ically not known to better than ∼100 K. Alto-
gether, the current level of systematic imperfec-
tions present in atmospheric models and observed
spectra of exoplanets (e.g., Greco & Brandt 2016)
mean that masses cannot yet be reliably measured
from fitting grids of synthetic spectra.
• Deuterium burning history. As brown dwarfs
with masses between about 13 MJup and 75 MJup
contract,
their core temperatures become hot
enough to burn deuterium, though not at suffi-
cient rates to balance surface radiative losses (e.g.,
12
)
r
y
G
(
e
g
A
10.000
1.000
0.100
0.010
0.001
Ross 458 C
HN Peg B
HD 203030 B
HR 8799 bcde
GU Psc b
51 Eri b
βPic b
2M0219−39 b
2M1207−39 b
HD 95086 b
ROXs42B b
2M0441+23 Bb
L0
L5
T0
T5
Y0
Spectral Type
Fig. 8.- Ultracool substellar companions with well-constrained ages and spectroscopically-derived classifications. Red circles are low-
mass stars and brown dwarfs (>13 MJup) while blue circles show companions near and below the deuterium-burning limit. Companions
are primarily from Deacon et al. (2014) together with additional discoveries from the literature.
Kumar 1963; Burrows & Liebert 1993).3 The onset
and timescale of deuterium burning varies primar-
ily with mass but also with metallicity, helium frac-
tion, and initial entropy (e.g., Spiegel et al. 2011);
lower-mass brown dwarfs take longer to initiate
deuterium burning than objects near the hydrogen-
burning minimum mass. This additional transient
energy source delays the otherwise invariable cool-
ing and causes luminosity tracks to overlap. Thus,
objects with the same luminosity and age can dif-
fer in mass depending on their deuterium-burning
history. Many substellar companions fall in this
ambiguous region, complicating mass determina-
tions by up to a factor of ∼2 (Figure 6 and Ta-
ble 1). With a large sample of objects in this region,
spectroscopy may ultimately be able to distinguish
higher- and lower-mass scenarios through relative
surface gravity measurements (Bowler et al. 2013).
• Planet composition. The gas and ice giants in
the Solar System are enriched in heavy elements
compared to solar values. The specific mecha-
nism for this enhancement is still under debate
but exoplanets formed via core accretion are ex-
pected to show similar compositional and abun-
dance ratio differences compared to their host stars,
whereas planets formed through cloud fragmenta-
tion or disk instability are probably quite similar
to the stars they orbit. The bulk composition of
planets modifies their atmospheric opacities and
influences both their emergent spectra and lumi-
nosity evolution (Fortney et al. 2008). A common
3 Solar-metallicity brown dwarfs with masses above ≈63 MJup
can also burn lithium. This limit changes slightly for non-solar
values (Burrows et al. 2001).
practice when deriving masses for imaged planets
is to assume solar abundances, which is largely dic-
tated by the availability of published atmospheric
and evolutionary models. Many of these assump-
tions can be removed with atmospheric retrieval
methods by directly fitting for atomic and molec-
ular abundances (Lee et al. 2013; Line et al. 2014;
Todorov et al. 2015).
• Additional sources of uncertainty. A num-
ber of other factors and implicit assumptions
can also introduce random and systematic un-
certainties in mass derivations. Different meth-
ods of PSF subtraction can bias photometry if
planet self-subtraction or speckle over-subtraction
is not properly corrected (e.g., Marois et al. 2006;
Lafreni`ere et al. 2007a;
Soummer et al. 2012).
Photometric variability from rapidly changing
or rotationally-modulated surface features can
introduce uncertainties in relative photometry
(e.g., Radigan et al. 2014; Metchev et al. 2015;
Zhou et al. 2016; Biller et al. 2015a). The host
star can also be variable if it has unusually large
starspot coverage or if it is very young and hap-
pens to have an edge-on disk (e.g., TWA 30A;
Looper et al. 2010).
Multiplicity can also bias luminosity measure-
ments. Roughly 20% of brown dwarfs are close bi-
naries with separation distributions peaking near
4.5 AU and mass ratios approaching unity (e.g.,
Burgasser et al. 2007; Kraus & Hillenbrand 2012;
Duchene & Kraus 2013).
If some planetary-mass
companions form in the same manner as brown
dwarfs, and if the same trends in multiplicity con-
tinue into the planetary regime, then a small frac-
tion of planetary-mass companions are probably
close, unresolved, equal-mass binaries. These sys-
tems will appear twice as luminous.
If atmospheric chemistry or cloud structure varies
latitudinally then orientation and viewing angle
could be important.
For the youngest proto-
planets embedded in their host stars' circumstel-
lar disks, accretion streams might dominate over
thermal photospheric emission, complicating lu-
minosity measurements and mass estimates (e.g.,
LkCa15 b and HD 100546 b; Kraus & Ireland 2012;
Quanz et al. 2013). Approaches to applying bolo-
metric corrections, measuring partly opaque coro-
nagraphs and neutral density filters, finely interpo-
lating atmospheric and evolutionary model grids,
or converting models between filter systems (for ex-
ample, CH4S to K) may vary. Finally, additional
energy sources like radioactivity or stellar insola-
tion are assumed to be negligible but could impact
the luminosity evolution of some exoplanets.
3.2. Measuring Masses
No imaged planet has yet had its mass measured. The
most robust, model-independent way to do so is through
dynamical interactions with other objects. Because plan-
ets follow mass-luminosity-age relationships, knowledge
of all three parameters are needed to test cooling models.
Once a mass is measured, its age (from the host star) and
bolometric luminosity (from its distance and spectral en-
ergy distribution) enable precision model tests, although
an assumption about energy losses from accretion via
hot-, warm-, or cold-start must be made. Nevertheless,
if all hot-start models overpredict the luminosities of gi-
ant planets, that would suggest that accretion history is
indeed an important factor in both planet formation and
realistic cooling models. Below is a summary of methods
to measure substellar masses.
• Dynamical masses. Most close-in (<100 AU)
planets have shown significant orbital motion since
their discoveries (Table 1). This relative motion
provides a measure of the total mass of the system
(Mstar+Mplanet).
If stationary background stars
can simultaneously be observed with the planet-
star pair then absolute astrometry is possible. This
then gives individual masses for each component
(Mstar and Mplanet separately). Unfortunately the
long orbital periods and lack of nearby background
stars for the present census of
imaged planets
means this method is currently impractical to mea-
sure masses.
Relative astrometry can also be combined with ra-
dial velocities to measure a planet's mass. Assum-
ing the visual orbit and total mass are well con-
strained from imaging, the mass of the companion
can be measured by monitoring the line of sight
reflex motion of the host star (e.g., Crepp et al.
2012a). This treats the system as a single-lined
spectroscopic binary, giving the mass function
m3
tot, where Mtot is the measured to-
tal mass, i is the measured inclination, and mp
is the mass of the planet.
If precise radial ve-
locities are not possible for the host star because
p sin3 i/M 2
13
it has an early spectral type (with few absorp-
tion lines) or high levels of stellar activity (RV
jitter) then RV monitoring of the planet can also
yield its mass. This can be achieved by combining
adaptive optics imaging and high-resolution near-
infrared spectroscopy to spatially separate the star
and planet, as has been demonstrated with β Pic b
(Snellen et al. 2014). Soon Gaia will produce pre-
cise astrometric measurements of the host stars of
imaged planets. Together with orbit monitoring
though high-contrast imaging, this may offer an-
other way to directly constrain the masses of im-
aged planets.
Close substellar binaries offer another approach.
Their orbital periods are typically faster and, in
rare cases when such binaries themselves orbit
a star, the age of the tertiary components can
be adopted from the host star. Several brown
dwarf-brown dwarf masses have been measured in
this fashion: HD 130948 BC (Dupuy et al. 2009),
Gl 417 BC (Dupuy et al. 2014), and preliminary
masses for ǫ Indi Bab (Cardoso et al. 2009). Iso-
lated substellar pairs are also useful for dynami-
cal mass measurements but their ages are gener-
ally poorly constrained unless they are members of
young clusters or moving groups. No binaries with
both components unambiguously residing in the
planetary-mass regime are known, but there is at
least one candidate (WISE J014656.66+423410.0;
Dupuy et al. 2015a).
• Keplerian disk rotation. Dynamical masses for
young protoplanets may eventually be possible us-
ing ALMA through Keplerian rotation of circum-
planetary disks. This requires resolving faint gas
emission lines (e.g., CO J=3–2 or CO J=2–1)
from the planet both spatially and spectrally, some-
thing that has yet to be achieved for known young
planets harboring subdisks (e.g., Isella et al. 2014;
Bowler et al. 2015a). Although challenging, this
type of measurement can act as a detailed probe of
the initial conditions of giant planet formation and
evolution.
• Stability analysis.
Numerical modeling of
planets and their interactions with debris disks,
protoplanetary disks, or additional planets offers
another way to constrain the mass of an imaged
planet.
If their masses are too low, planets will
not be able to gravitationally shape dust and
planetesimals in a manner consistent with obser-
vations. Modeling of the disks and companions
orbiting Fomalhaut and β Pic illustrate this
approach;
independent constraints on the orbit
and masses of these companions can be made by
combining spatially-resolved disk structures (a
truncated, offset dust ring encircling Fomalhaut
and a warped inner disk surrounding β Pic b)
and observed orbital motion (e.g., Chiang et al.
2009; Dawson et al. 2011; Kalas et al. 2013;
Beust et al. 2014; Millar-Blanchaer et al. 2015).
Likewise,
if a planet's mass is too high then it
carves a larger disk gap or may destabilize other
planets in the system through mutual interactions.
14
For example, detailed N -body simulations of
HR 8799's planets have shown that they must
have masses .10–20 MJup - consistent with giant
planet evolutionary models - or they would have
become dynamically unstable by the age of the
(e.g., Go´zdziewski & Migaszewski
host
2009;
Fabrycky & Murray-Clay
2010;
Currie et al. 2011b; Sudol & Haghighipour 2012;
Go´zdziewski & Migaszewski 2014).
star
• Disk morphology. Large-scale structures in disks
- clumps, asymmetries, warps, gaps, rings, trun-
cated edges, spiral arms, and geometric offsets -
can also be used to indirectly infer the presence
of unseen planets and predict their masses and
locations (e.g., Wyatt et al. 1999; Ozernoy et al.
2000; Kenyon & Bromley 2004). This approach
relies on assumptions about disk surface density
profiles and grain properties, so it is not a com-
pletely model-free measurement, but it is poten-
tially sensitive to planet masses as low as a few
tens of Earth masses (Rosotti et al. 2016).
It
also enables an immediate mass evaluation with-
out the need for long-term orbit monitoring. Re-
cently, Dong et al. (2015) and Zhu et al. (2015)
presented a novel approach along these lines to
predict the locations and masses widely-separated
companions inducing spiral arms on a circumstellar
disk (see also Dong et al. 2016b, Dong et al. 2016a,
and Jilkova & Zwart 2015). This may prove to be
a valuable way to constrain masses of planetary
companions at extremely wide orbital distances.
4. SURVEY OF SURVEYS
Myriad large high-contrast imaging surveys have been
carried out over the past decade4. The most impact-
ful programs are highly focused, carefully designed with
well-understood biases, and have meticulously-selected
target lists to address specific science questions. The
advantages of large surveys include homogeneous obser-
vations, instrument setups, data reduction pipelines, and
statistical treatments of the results.
Below are summaries of the most substantial high-
contrast imaging surveys carried out to date with a focus
on deep adaptive optics imaging programs that routinely
reach planetary masses and employ modern observing
and post-processing techniques to suppress speckle noise.
These surveys produced the first wave of discoveries (Fig-
ure 9), opening the door to directly characterizing the
atmospheres of exoplanets as well as their orbits through
astrometric monitoring (Figure 10 and Appendix 4) This
section follows an historical approach by outlining early
ground- and space-based experiments, the first gener-
ation of planet-finding instruments and associated sur-
veys, and the next generation of instruments character-
ized by extreme adaptive optics systems with exception-
ally high Strehl ratios.
4.1. Early Surveys
4 The basic properties for most of these programs until 2014 are
summarized in Table 1 of Chauvin et al. (2015).
image
imaging
surveys
Early high-contrast
in search
of closely-separated brown dwarf companions and
giant planets were conducted with speckle interfer-
ometry (Henry & McCarthy 1990),
stabiliz-
(Nakajima et al. 1994), HST (Sartoretti et al.
ers
1998;
Schroeder et al. 2000; Brandner et al. 2000;
Lowrance et al. 2005; Luhman et al. 2005), speckle cam-
eras (Neuhauser et al. 2003), or newly-commissioned
adaptive optics systems from the ground with facility
instruments (Oppenheimer et al. 2001; Macintosh et al.
2001; Chauvin et al. 2003; McCarthy & Zuckerman
2004;
2005;
Tanner et al. 2007). When PSF subtraction was
performed, it usually entailed roll-subtraction (for HST ;
e.g., Liu 2004), self-subtraction with a rotated PSF, or
reference star subtraction.
Nakajima et al.
Carson et al.
2005;
Some of these pioneering programs are especially
noteworthy for their depth and emphasis on statisti-
cal results.
Lowrance et al. (2005) targeted 45 sin-
gle young A–M stars with NICMOS in the F 160W
filter (1.6 µm) on board HST . Two brown dwarfs
were uncovered, TWA 5 B (Lowrance et al. 1999) and
HR 7372 B (Lowrance et al. 2000), as well as Gl 577 BC,
a tight binary companion near the hydrogen-burning
limit. Masciadri et al. (2005) used NaCo at the VLT
to obtain deep adaptive optics imaging of 28 young
nearby stars. No substellar companions were found, but
the importance of thoroughly reporting survey results
is highlighted, even for non-detections, a theme that
continues today. Focusing exclusively on young moving
group members in L′-band enabled Kasper et al. (2007)
to reach exceptionally low limiting masses for a sample
of 22 stars with NaCo. The Palomar and Keck adap-
tive optics survey by Metchev & Hillenbrand (2009) is
another especially valuable contribution; they imaged
266 FGK stars and discovered two brown dwarf com-
panions, HD 49197 B (Metchev & Hillenbrand 2004) and
HD 203030 B (Metchev & Hillenbrand 2006), implying a
substellar occurrence rate of 3.2+3.1
−2.7%. HD 203030 B
was the first young brown dwarf for which signs of a dis-
crepancy between the field spectral type-effective tem-
perature sequence was recognized, now understood as a
retention of clouds to lower effective temperatures at low
surface gravities. These groundbreaking surveys helped
define the scientific motivation, framework, and early ex-
pectations for the first generation planet-finding instru-
ments and larger observing programs.
4.2. The First Generation: Dedicated Instruments,
Expansive Surveys, and Innovative Speckle
Suppression Techniques
High-contrast imaging is largely driven by advances
in instrumentation and speckle suppression. The first
wave of instruments specifically designed to image giant
planets gave rise to large surveys (N ≈50–500) targeting
mostly young nearby stars.
Deep observations in
pupil-tracking mode (angular differential imaging) have
become standardized as a way to distinguish quasi-static
speckles from planets (Marois et al. 2006). This era
is also characterized by the advent of advanced PSF
subtraction techniques to optimally remove speckles
during post-processing.
Two especially important
algorithms are the Locally Optimized Combination
b
HR 8799
c
e
d
Maire et al. (2015)
LBT/LMIRCam L′
Discovery: Marois et al. (2008, 2010)
E
1′′
39 AU
HD 95086
b
Galicher et al. (2014)
Gemini−S/GPI K1
Discovery: Rameau et al. (2013a, 2013b)
E
0.3′′
27 AU
N
N
β Pic
0.5′′
10 AU
b
Nielsen et al. (2014)
Gemini−S/NICI KCont
Discovery: Lagrange et al. (2009, 2010)
E
2′′
39 AU
51 Eri
15
N
b
N
E
Macintosh et al. (2015)
Gemini−S/GPI H
0.5′′
15 AU
Fig. 9.- Gallery of imaged planets at small separations (<100 AU). HR 8799 harbors four massive planets (5–10 MJup) at orbital
distances of 15–70 AU (Marois et al. 2008; Marois et al. 2010b), β Pic hosts a nearly edge-on debris disk and a ≈13 MJup planet at
9 AU (Lagrange et al. 2009a; Lagrange et al. 2010), a ≈5 MJup planet orbits HD 95086 at 56 AU (Rameau et al. 2013c; Rameau et al.
2013b), and 51 Eri hosts a ∼2 MJup planet at 13 AU (Macintosh et al. 2015). Images are from Maire et al. (2015a), Nielsen et al. (2014),
Galicher et al. (2014), and Macintosh et al. (2015).
of Images (LOCI; Lafreni`ere et al. 2007a), which is
based on least-squares minimization of residual speckle
noise, and Karhunen-Lo`eve Image Projection (KLIP;
Soummer et al. 2012), a computationally-fast method
based on principal component analysis. The introduc-
tion of these new methods gave rise to an array of
sophisticated data reduction pipelines with additional
features aimed at minimizing biases and avoiding both
self- and over-subtraction of planet flux in ADI and
SDI datasets (Marois et al. 2010a; Amara & Quanz
2012;
2013a;
Wahhaj et al. 2013a; Brandt et al. 2013; Fergus et al.
2014; Mawet et al. 2014; Marois et al. 2014; Currie et al.
2014c; Cantalloube et al. 2015; Rameau et al. 2015;
Wahhaj et al. 2015; Savransky 2015; Dou et al. 2015;
Hagelberg et al. 2016; Gonzalez et al. 2016).
2012; Meshkat et al.
Pueyo et al.
The suite of instrumentation for high-contrast imag-
ing has ballooned over the past 15 years and in-
cludes dual-channel imagers, infrared wavefront sensors,
non-redundant aperture masking interferometry, adap-
tive secondary mirrors, integral field units, high-order
adaptive optics systems, and specialized coronagraphs
(e.g., apodized Lyot coronagraph, annular groove phase
mask coronagraph, vector vortex coronagraph, apodizing
phase plate, and four quadrant phase mask; Rouan et al.
2000; Guyon et al. 2005; Soummer 2005; Mawet et al.
2005; Kenworthy et al. 2007; Mawet et al. 2010). Many
of these have been implemented in the first generation of
instruments in part as testbeds for regular use in second-
generation systems. These instruments are reviewed
in detail
in Guyon et al. (2006), Beuzit et al. (2007),
Oppenheimer & Hinkley (2009), Perryman (2011), and
Mawet et al. (2012b).
4.2.1. VLT and MMT Simultaneous Differential Imager
Survey
This survey (PI: B Biller) targeted 45 young stars
between 2003–2006 with ages .250 Myr and distances
b
2020
2030
2040
2050
16
)
"
(
c
e
D
∆
2
1
0
−1
−2
2
HR 8799
c
e
d
2026
2023
β Pic
2020
2017
)
"
(
c
e
D
∆
0.4
0.2
0.0
−0.2
−0.4
1
0
∆RA (")
−1
−2
0.4
0.2
0.0
∆RA (")
−0.2
−0.4
Fig. 10.- The orbits of HR 8799 bcde and β Pic b. Astrometric monitoring has revealed significant orbital motion over the past decade,
enabling measurements of their Keplerian orbital parameters and offering clues about their dynamical history. The HR 8799 planets are
on low-eccentricity orbits with some evidence that HR 8799 d may be mutually misaligned with the other planets, which otherwise appear
to be coplanar (e.g., Pueyo et al. 2015). β Pic b follows a nearly edge-on orbit with a low eccentricity (e.g., Millar-Blanchaer et al. 2015).
Astrometric measurements are compiled in Appendix 4. Orbits depicted here are from Zurlo et al. (2016) and Nielsen et al. (2014).
within 50 pc using Simultaneous Differential Imagers
mounted on the VLT and MMT (Biller et al. 2007). It
was among the first to utilize simultaneous differential
imaging to search for cold planets around a large sample
of young stars. The SDI method takes advantage of ex-
pected spectral differences between the star, which has a
nearly flat continuum, and cool, methanated planets by
simultaneously imaging in multiple narrow-band filters
across this deep absorption feature at 1.6 µm. Because
speckles radially scale with wavelength while real objects
remain stationary, their observations also had some sen-
sitivity to warmer planets without methane (though it
is now clear that the onset of methane occurs at lower
temperatures for giant planets than for brown dwarfs).
No substellar companions were found, which ruled out
a linearly-flat extension of close-in giant planets out to
45 AU with high confidence.
4.2.2. GDPS: Gemini Deep Planet Survey
GDPS (PI: D. Lafreni`ere) was a large high-contrast
imaging program at the Gemini-North 8.1-m telescope
with the NIRI camera and Altair AO system focusing on
85 stars, 16 of which were identified as close multiples
(Lafreni`ere et al. 2007b). The sample contained a mix of
nearby GKM stars within 35 pc comprising then-known
or suspected nearby young moving group members, stars
with statistically young ages, and several others harbor-
ing circumstellar disks. Altogether the ages span 10 Myr
to ∼5 Gyr. The observations were taken in ADI mode
with the CH4S filter, and PSF subtraction was carried
out with the LOCI algorithm (Lafreni`ere et al. 2007a).
No substellar companions were discovered, implying an
occurrence rate of <23% for >2 MJup planets between
25–420 AU and <12% for >2 MJup planets between 50–
295 AU at the 95% confidence level.
4.2.3. MMT L′ and M -Band Survey of Nearby Sun-Like
Stars
Heinze et al. (2010b) carried out a deep L′- and M -
band survey of 54 nearby FGK stars at the MMT with
Clio. The MMT adaptive optics system uses a de-
formable secondary mirror which reduces the thermal
background by minimizing the number of optical ele-
ments along the light path. Observations were car-
ried out between 2006–2007 with angular differential
imaging. The image processing pipeline is described
in Heinze et al. (2008) and Heinze et al. (2010b). The
target ages are generally older (∼0.1–2 Gyr) but the
long wavelengths of the observations and proximity of
the sample (.25 pc) enabled sensitivity to planetary
masses for most of the targets. One new low-mass stellar
companion was discovered and the binary brown dwarf
HD 130948 BC was recovered in the survey. The statisti-
cal results are detailed in Heinze et al. (2010a); they find
that no more than 50% of Sun-like stars host ≥5 MJup
planets between 30–94 AU and no more than 15% host
≥10 MJup planets between 22–100 AU at the 90% confi-
dence level.
4.2.4. NaCo Survey of Young Nearby Austral Stars
This program utilized NaCo at the VLT between 2002–
2007 to target 88 young GKM stars within 100 pc
(Chauvin et al. 2010). 17 new close multiple systems
were uncovered and deep imaging was obtained for 65
single young stars. Observations were taken with a
Lyot coronagraph in H and KS bands and PSF sub-
traction was performed with azimuthally-averaged sub-
traction and high-pass filtering.
The most important discovery from this survey was
2M1207–3932 b, a remarkable 5 ± 2 MJup companion to
a 25 MJup brown dwarf in the 10 Myr TWA moving group
(Chauvin et al. 2004; Chauvin et al. 2005a) enabled with
infrared wavefront sensing. The unusually red colors
and spectral shape of 2M1207–3932 b (Patience et al.
2010) have made it the prototype of young dusty L
dwarfs, now understood as a cloudy extension of the
L dwarf sequence to low temperatures (Barman et al.
2011b; Marley et al. 2012). This system is also unusual
from the perspective of brown dwarf demographics; the
mass ratio of ∼0.2 and separation of ∼41 AU make it an
outlier compared to brown dwarf mass ratio and sepa-
ration distributions in the field (Burgasser et al. 2007).
Two other substellar companions were discovered in this
survey: GSC 08047-00232 B (Chauvin et al. 2005c), also
independently found by Neuhauser et al. (2003), and
AB Pic B (Chauvin et al. 2005b), which resides near the
deuterium-burning limit.
4.2.5. NaCo Survey of Young Nearby Dusty Stars
This VLT/NaCo survey targeted 59 young nearby
AFGK stars with ages .200 Myr and distances within
65 pc (Rameau et al. 2013a). Most of the sample are
members of young moving groups and the majority (76%)
were chosen to have mid-infrared excesses, preferentially
selected for having debris disks. Observations were car-
ried out in L′-band between 2009–2012 using angular dif-
ferential imaging. Four targets in the sample had known
substellar companions (HR 7329, AB Pic, HR 8799, and
β Pic). No new substellar companions were discovered
but eight new visual binaries were resolved. A statistical
analysis of AF stars between 5–320 AU and 3–14 MJup
implies a giant planet occurrence rate of 7.4+3.6
−2.4% (68%
confidence level).
4.2.6. SEEDS: Strategic Exploration of Exoplanets and
Disks with Subaru
The SEEDS survey (PI: M. Tamura) was a 125-night
program on the 8.2-m Subaru Telescope targeting about
500 stars to search for giant planets and spatially re-
solve circumstellar disks (Tamura et al. 2009). Tamura
(2016) provide an overview of the observing strategy,
target samples, and main results. Observations were
carried out with the HiCIAO camera behind Subaru's
AO188 adaptive optics system over five years beginning
in 2009. The sample contained a mixture of young stars
in star-forming regions, moving groups, and open clus-
ters; nearby stars and white dwarfs; and stars with pro-
toplanetary disks and debris disks. Most of the observa-
tions were taken in H-band in angular differential imag-
ing mode as well as polarimetric differential imaging for
young disk-bearing stars. The ADI reduction pipeline is
described in Brandt et al. (2013).
Three new substellar companions were found in
SEEDS: GJ 758 B (Thalmann et al. 2009), κ And B
(Carson et al. 2013), and GJ 504 b (Kuzuhara et al.
2013). The masses of GJ 504 b and κ And B may
fall in the planetary regime depending on the ages and
metallicities of the system, which are still under debate.
Two brown dwarf companions found in the Pleiades
(HD 23514 B and HII 1348 B; Yamamoto et al. 2013)
had also independently been discovered by other groups.
SEEDS resolved a remarkable number of protoplanetary
and transition disks in polarized light- over two dozen
in total - revealing previously unknown gaps, rings,
17
and spiral structures down to 0.′′1 with exceptional clar-
ity (e.g., Thalmann et al. 2010; Hashimoto et al. 2011;
Mayama et al. 2012; Muto et al. 2012).
The statistical results for debris disks are presented in
Janson et al. (2013c). At the 95% confidence level, they
find that <15–30% of stars host >10 MJup planets at
the gap edge. Brandt et al. (2014a) inferred a frequency
of 1.0–3.1% at the 68% confidence level for 5–70 MJup
companions between 10–100 AU by combining results
from the SEEDS moving group sample (Brandt et al.
2014c), the SEEDS disk sample (Janson et al. 2013c),
the SEEDS Pleiades sample (Yamamoto et al. 2013),
GDPS (Lafreni`ere et al. 2007b), and the NICI Campaign
moving group sample (Biller et al. 2013).
4.2.7. Gemini NICI Planet-Finding Campaign
The Gemini NICI Planet-Finding Campaign (PI: M.
Liu) was a 500-hour survey targeting about 230 young
stars of all spectral classes with deep imaging using
the Near-Infrared Coronagraphic Imager on the Gemini-
South 8.1-m telescope (Liu et al. 2010). NICI is an imag-
ing instrument encompassing an adaptive optics system,
tapered and partly-translucent Lyot coronagraph, and
dual-channel camera (Chun et al. 2008). Campaign ob-
servations spanned 2008–2012 and were carried out in
two modes: single-channel H-band with angular differ-
ential imaging, and simultaneous dual-channel (CH4S at
1.578 µm and CH4L at 1.652 µm) angular and spectral
differential imaging to maximize sensitivity to methane-
dominated planets. The observing strategy and re-
duction pipeline are detailed in Biller et al. (2008) and
Wahhaj et al. (2013a), and NICI astrometric calibration
is discussed in Hayward et al. (2014).
One previously-known brown dwarf companion was re-
solved into a close binary, HIP 79797 Bab (Nielsen et al.
2013), and three new substellar companions were found:
PZ Tel B, a highly eccentric brown dwarf companion
in the β Pic moving group (Biller et al. 2010); CD-
35 2722 B, a young mid-L dwarf in the AB Dor moving
group (Wahhaj et al. 2011); and HD 1160 B, a substellar
companion orbiting a young massive star (Nielsen et al.
2012). No new planets were discovered but β Pic b was
recovered during the survey and its orbit was shown to be
misaligned with the inner and outer disks (Nielsen et al.
2014; Males et al. 2014). Two debris disks surround-
ing HR 4796 A and HD 141569 were also resolved with
unprecedented detail (Wahhaj et al. 2014; Biller et al.
2015b).
The statistical results are organized in several stud-
ies. From a sample of 80 members of young moving
groups, Biller et al. (2013) measured the frequency of 1–
20 MJup planets between 10–150 AU to be <6–18% at
the 95.4% confidence level, depending on which hot-start
evolutionary models are adopted. The high-mass sample
of 70 B and A-type stars was described in Nielsen et al.
(2013); they found that the frequency of >4 MJup plan-
ets between 59–460 AU is <20% at 95% confidence.
Wahhaj et al. (2013b) found that <13% of debris disk
stars have ≥5 MJup planets beyond 80 AU at 95% con-
fidence from observations of 57 targets.
4.2.8. IDPS: International Deep Planet Search
IDPS is an expansive imaging survey carried out at
the VLT with NaCo, Keck with NIRC2, Gemini-South
18
with NICI, and Gemini-North with NIRI targeting ≈300
young A–M stars (PI: C. Marois). This 14-year survey
was mostly carried out in K band, though much of the
survey comprised a mix of broad- and narrow-band near-
infrared filters. Target ages were mostly .300 Myr and
encompassed distances from ∼10–80 pc (Galicher et al.
2016, submitted).
The main result from this survey was the discovery of
the HR 8799 planets (Marois et al. 2008; Marois et al.
2010b). Altogether over 1000 unique point sources were
found, most of which were meticulously shown to be
background stars from multi-epoch astrometry (Galicher
et al. 2016, submitted). The preliminary analysis of
a subset of high-mass A and F stars spanning ≈1.5–
3.0 M⊙ was presented in Vigan et al. (2012). 39 new
observations in H, K, and CH4S filters were carried out
in angular differential imaging mode between 2007–2012
and were combined with three high-mass targets from
the literature. Stellar ages span 8–400 Myr with dis-
tances out to 90 pc and comprise a mix of young moving
group members, young field stars, and debris disk hosts.
The subsample of 42 massive stars includes three hosts
of substellar companions: HR 8799, β Pic, and HR 7329,
a β Pic moving group member with a wide brown dwarf
companion (Lowrance et al. 2000). Including the detec-
tions of planets around HR 8799 and β Pic, Vigan et al.
(2012) measure the occurrence rate of 3–14 MJup planets
between 5–320 AU to be 8.7+10.1
−2.8 % at 68% confidence.
The complete statistical analysis for the entire sample
is presented in Galicher et al. (2016, submitted). They
merge their own results for 292 stars with the GDPS
and NaCo-LP surveys, totaling a combined sample of
356 targets. From this they infer an occurrence rate
of 1.05+2.80
−0.70% (95% confidence interval) for 0.5–14 MJup
planets between 20–300 AU. They do not find evidence
that this frequency depends on stellar host mass. In ad-
dition, 16 of the 59 binaries resolved in IDPS are new.
4.2.9. PALMS: Planets Around Low-Mass Stars
The PALMS survey (PI: B. Bowler) is a deep imaging
search for planets and brown dwarfs orbiting low-mass
stars (0.1–0.6 M⊙) carried out at Keck Observatory with
NIRC2 and Subaru Telescope with HiCIAO. Deep coron-
agraphic observations were acquired for 78 single young
M dwarfs in H- and Ks-bands between 2010–2013 us-
ing angular differential imaging. An additional 27 stars
were found to be close binaries. Targets largely originate
from Shkolnik et al. (2009), Shkolnik et al. (2012), and
an additional GALEX-selected sample (E. Shkolnik et
al., in preparation). Most of these lie within 40 pc and
have ages between 20–620 Myr; about one third of the
sample are members of young moving groups. The ob-
servations and PSF subtraction pipeline are described in
Bowler et al. (2015b).
3629
B
GJ
Four substellar companions were found in this pro-
1RXS J235133.3+312720 B (Bowler et al.
gram:
2012b),
2012a),
2015b),
1RXS J034231.8+121622 B (Bowler et al.
and 2MASS J15594729+4403595 B (Bowler et al.
2015b).
1RXS J235133.3+312720 B is a particularly
useful benchmark brown dwarf because it orbits a mem-
ber of a young moving group (AB Dor) and therefore
has a well-constrained age (≈120 Myr).
(Bowler et al.
The statistical results from the survey are presented in
Bowler et al. (2015b). No planets were found, implying
an occurrence rate of <10.3% for 1–13 MJup planets be-
tween 10–100 AU at the 95% confidence level assuming
hot-start models and <16.0% assuming cold-start mod-
els. For the most massive planets between 5–13 MJup,
the upper limits are <6.0% and <9.9% for hot- and cold-
start cooling models.
The second, parallel phase of the PALMS survey is an
ongoing program targeting a larger sample of ∼400 young
M dwarfs primarily at Keck with shallower contrasts
(Bowler et al., in prep.). Initial discoveries include two
substellar companions: 2MASS J01225093–2439505 B,
an L-type member of AB Dor at the planet/brown dwarf
boundary with an unusually red spectrum (Bowler et al.
2013; Hinkley et al. 2015a), and 2MASS J02155892–
0929121 C (Bowler et al. 2015c), a brown dwarf in a close
quadruple system which probably belongs to the Tuc-Hor
moving group.
4.2.10. NaCo-LP: VLT Large Program to Probe the
Occurrence of Exoplanets and Brown Dwarfs at
Wide Orbits
The NaCo-LP survey was a Large Program at the VLT
focused on 86 young, bright, primarily FGK stars (PI: J.-
L. Beuzit). H-band observations were carried out with
NaCo in ADI mode between 2009–2013 (Chauvin et al.
2015). The target sample is described in detail
in
Desidera et al. (2015); stars were chosen to be single,
have ages .200 Myr, and lie within 100 pc. Many of
these stars were identified as new members of young mov-
ing groups.
Although no new substellar companions were dis-
covered, an intriguing white dwarf was found orbiting
HD 8049, an ostensibly young K2 star that may instead
be much older due to mass exchange with its now evolved
companion (Zurlo et al. 2013). New observations of the
spatially resolved debris disk around HD 61005 ("the
Moth") were presented by Buenzli et al. (2010), and 11
new close binaries were resolved during this program
(Chauvin et al. 2015).
The statistical analysis of the sample of single stars
was performed in Chauvin et al. (2015). Based on a sub-
sample of 51 young FGK stars, they found that <15% of
Sun-like stars host planets with masses >5 MJup between
100–500 AU and <10% host >10 MJup planets between
50–500 AU at the 95% confidence level. Reggiani et al.
(2016) use these NaCo-LP null results together with ad-
ditional deep archival observations to study the compan-
ion mass function as it relates to binary star formation
and planet formation. From their full sample of 199 Sun-
like stars, they find that the results from direct imag-
ing are consistent with the superposition of the planet
mass function determined from radial velocity surveys
and the stellar companion mass ratio distribution down
to 5 MJup, suggesting that many planetary-mass com-
panions uncovered with direct imaging may originate
from the tail of the brown dwarf mass distribution in-
stead of being the most massive representatives of the
giant planet population.
4.3. Other First Generation Surveys
Several smaller, more focused surveys have also been
carried out with angular differential imaging: Apai et al.
2008 targeted 8 debris disk hosts with NaCo us-
ing simultaneous differential imaging; Ehrenreich et al.
(2010) imaged 38 high-mass stars primarily with NaCo;
Janson et al. (2011a) observed 15 B and A stars with
NIRI; Delorme et al. (2012) presented observations of
16 young M dwarfs in L′-band with NaCo; Maire et al.
(2014) targeted 16 young AFGK stars with NaCo's four-
quadrant phase mask, simultaneous differential imaging,
and angular differential imaging; Meshkat et al. (2015a)
used NaCo's Apodizing Phase Plate coronagraph in L′
band to image six young debris disk hosts with gaps as
part of the Holey Debris Disk survey; and Meshkat et al.
(2015b) also used the APP at NaCo to image a sam-
ple of 13 A- and F-type main sequence stars in search
of planets, uncovering a probable brown dwarf around
HD 984 (Meshkat et al. 2015c). The Lyot Project is an-
other survey with important contributions for its early
use of coronagraphy behind an extreme adaptive op-
tics system (Oppenheimer et al. 2004). This survey
was carried out at the 3.6-m AEOS telescope equipped
with a 941-actuator deformable mirror and targeted 86
nearby bright stars using angular differential imaging
(Leconte et al. 2010).
4.4. The Second Generation: Extreme Adaptive Optics,
Exceptional Strehl Ratios, and Optimized Integral
Field Units
The transition to second-generation planet-finding in-
struments began over the past few years. This new era
is characterized by regular implementation of high-order
("extreme") adaptive optics systems with thousands of
actuators and exceptionally low residual wavefront er-
rors; pyramid wavefront sensors providing better sensi-
tivity and higher precision wavefront correction; Strehl
ratios approaching (and often exceeding) 90% at near-
infrared wavelengths; high-contrast integral field units
designed for on-axis observations enabling speckle sub-
traction and low-resolution spectroscopy; sensitivity to
smaller inner working angles than first-generation instru-
ments; and advanced coronagraphy.
19
Results from the Project 1640 survey include several
discoveries of faint stellar companions to massive A stars
(Zimmerman et al. 2010; Hinkley et al. 2010) and follow-
up astrometric and spectral characterization of known
substellar companions (Crepp et al. 2015; Hinkley et al.
2013).
In addition, Oppenheimer et al. (2013) and
Pueyo et al. (2015) presented detailed spectroscopic and
astrometric analysis of the HR 8799 planets and found
intriguing evidence for mutually dissimilar spectral prop-
erties and signs of non-coplanar orbits.
4.4.2. LEECH: LBTI Exozodi Exoplanet Common Hunt
LEECH is an ongoing ∼70-night high-contrast imag-
ing program (PI: A. Skemer) at the twin 8.4-m Large
Binocular Telescope. Survey observations began in 2013
and are carried out in angular differential imaging mode
in L′-band with LMIRcam utilizing deformable sec-
ondary mirrors to maximize sensitivity at mid-IR wave-
lengths by limiting thermal emissivity from warm op-
tics (Skemer et al. 2014b). The target sample focuses on
intermediate-age stars <1 Gyr including members of the
∼500 Myr Ursa Majoris moving group, massive BA stars,
and nearby young FGK stars.
In addition to searching for new companions, this sur-
vey is also characterizing known planets using the unique
mid-IR instrumentation, sensitivity, and filter suite at
the LBT. Maire et al. (2015b) refined the orbits of the
HR 8799 planets and found them to be consistent with
8:4:2:1 mean motion resonances. Skemer et al. (2016)
observed GJ 504 b in three narrow-band filters spanning
3.7–4.0 µm. Model fits indicate an exceptionally low
effective temperature of ≈540 K and enhanced metal-
licity, possibly pointing to an origin through core ac-
cretion. Additionally, Schlieder et al. (2016) presented
LEECH observations and dynamical mass measurements
the Ursa Majoris binary NO UMa. Recently the integral
field unit Arizona Lenslets for Exoplanet Spectroscopy
(ALES; Skemer et al. 2015) was installed inside LMIR-
cam and will enable integral-field spectroscopy of planets
between 3–5 µm for the first time.
4.4.1. Project 1640
4.4.3. GPIES: Gemini Planet Imager Exoplanet Survey
Project 1640 is a large ongoing survey (PI: R. Op-
penheimer) and high-contrast imaging instrument with
the same name located behind the PALM-3000 second-
generation adaptive optics system at the Palomar Obser-
vatory 200-inch (5.1-meter) Hale Telescope. The instru-
ment itself contains an apodized-pupil Lyot coronagraph
and integral field unit that samples 32 spectral channels
across Y , J, and H bands (Hinkley et al. 2011), produc-
ing a low-resolution spectrum to broadly characterize the
physical properties of faint companions (Roberts et al.
2012; Rice et al. 2015). The survey consists of two
phases, the first (now concluded) with the original Palo-
mar Adaptive Optics System (PALM-241; Troy et al.
2000) and a second ongoing three-year program focus-
ing on nearby massive stars with the upgraded PALM-
3000 adaptive optics system (Dekany et al. 2013). The
data reduction pipeline is described in Zimmerman et al.
(2011) and a detailed treatment of speckle subtrac-
tion, precision astrometry, and robust spectrophotom-
etry can be found in Crepp et al. (2011), Pueyo et al.
(2012), Oppenheimer et al. (2013), Fergus et al. (2014),
and Pueyo et al. (2015).
GPIES is an ongoing 890-hour, 600-star survey to im-
age extrasolar giant planets and debris disks with the
Gemini Planet Imager at Gemini-South (PI: B. Mac-
intosh). GPI is expressly built to image planets at
small inner working angles; its high-order adaptive op-
tics system incorporates an apodized pupil Lyot coro-
nagraph, integral field spectrograph, imaging polarime-
ter, and (imminent) non-redundant masking capabilities
(Macintosh et al. 2014). Survey observations targeting
young nearby stars began in 2014 and will span three
years.
Macintosh et al. (2015) presented the discovery of 51
Eri b, the first exoplanet found in GPIES and the lowest-
mass planet imaged in thermal emission to date. This re-
markable young, methanated T dwarf has a contrast of
14.5 mag in H-band at a separation of 0.′′45, which trans-
lates into a mass of only 2 MJup at 13 AU assuming hot-
start cooling models. It is also the only imaged planet
consistent with the most pessimistic cold-start evolution-
ary models, in which case its mass may be as high as
12 MJup. De Rosa et al. (2015) obtained follow-up obser-
vations with GPI and showed that 51 Eri b shares a com-
20
mon proper motion with its host and exhibits slight (but
significant) orbital motion. Other initial results from this
survey include astrometry and a refined orbit for β Pic b
(Macintosh et al. 2014), as well as resolved imaging of
the debris disks around HD 106906 (Kalas et al. 2015)
and HD 131835 (Hung et al. 2015).
4.4.4. SPHERE GTO Survey
SPHERE (Spectro-Polarimetric High-contrast Exo-
planet Research) is an extreme adaptive optics system
(SAXO) and versatile instrument for high-contrast imag-
ing and spectroscopy at the VLT with a broad range of
capabilities (Beuzit et al. 2008).
IRDIS (Dohlen et al.
2008) offers classical and dual-band imaging (Vigan et al.
2010), dual-polarization imaging (Langlois et al. 2014),
and long-slit spectroscopy (Vigan et al. 2008);
IFS
provides low-resolution (R∼30–50) integral field spec-
troscopy spanning 0.95–1.65 µm (Claudi et al. 2008);
and ZIMPOL enables diffraction-limited imaging and po-
larimetry in the optical (Thalmann et al. 2008). As part
of a guaranteed time observing program, the SPHERE
GTO team is carrying out a large ongoing survey (PI:
J.-L. Beuzit) with a range of science goals. About 200
nights of this time are devoted to a deep near-infrared
imaging survey (SHINE: SpHere Infrared survEy) to
search for and characterize exoplanets around 400–600
stars, while the remaining ∼60 nights will be used for
a broad range of science including spatially-resolved ob-
servations of circumstellar disks and optical imaging of
planets.
Initial results from the SHINE survey include ob-
servations of the brown dwarf GJ 758 B (Vigan et al.
2016), PZ Tel B and HD 1160 B (Maire et al. 2016),
HD 106906 (Lagrange et al. 2016), and the HR 8799
planets (Zurlo et al. 2016; Bonnefoy et al. 2016). Other
results have focused on resolved imaging of the debris
disk surrounding HD 61005, which may be a product of
a recent planetesimal collision (Olofsson et al. 2016), and
HD 135344 B, host of a transition disk with striking spi-
ral arm structure (Stolker et al. 2016). Boccaletti et al.
(2015) uncovered intriguing and temporally evolving fea-
tures in AU Mic's debris disk. In addition, Garufi et al.
(2016) presented deep IRDIS near-infrared images and
visible ZIMPOL polarimetric observations of HD 100546
revealing a complex disk environment with considerable
structure and resolved K-band emission at the location
of the candidate protoplanet HD 100546 b.
4.4.5. Other Second Generation Instruments and Surveys
A number of other novel instruments and forthcom-
ing surveys bear highlighting. MagAO (PI: L. Close) at
the Magellan 6.5-m Clay telescope is a versatile adap-
tive optics system consisting of a 585-actuator adap-
tive secondary mirror, pyramid wavefront sensor, and
two science cameras offering simultaneous diffraction-
limited imaging spanning the visible (0.6–1.05 µm)
with VisAO and near-infrared (1–5.3 µm) with Clio2
(Close et al. 2012; Morzinski et al. 2014; Morzinski et al.
2015). Strehl ratios of ∼20–30% in the optical are open-
ing up new science fronts including deep red-optical ob-
servations of exoplanets (Males et al. 2014; Wu et al.
2015), characterization of accreting protoplanets in Hα
(Sallum et al. 2015a), and high spatial resolution imag-
ing down to ∼20 mas (Close et al. 2013). Vector apodiz-
ing phase plate coronagraphs were recently installed in
MagAO and other upgrades such as an optical integral
field unit are possible in the future.
Subaru Coronagraphic Extreme Adaptive Optics
(SCExAO; PI: O. Guyon) is being built for the Subaru
telescope and is the newest extreme adaptive optics sys-
tem on a large telescope. A detailed description of all
facets of this instrument is described in Jovanovic et al.
(2015). In short, a pyramid wavefront sensor is coupled
with a 2000-element deformable mirror to produce Strehl
ratios in excess of 90%. The instrument is particularly
flexible, allowing for a variety of setups and instrument
subcomponents including speckle nulling to suppress
static and slowly changing speckles (Martinache et al.
2014), a near-infrared science camera (currently Hi-
CIAO), sub-diffraction-limited interferometric science in
the visible with VAMPIRES (Norris et al. 2015) and
FIRST (Huby et al. 2012), high-contrast integral field
spectroscopy (Brandt et al. 2014b), and coronagraphy
with phase-induced amplitude apodization (Guyon 2003)
and vector vortex coronagraphs (Mawet et al. 2010).
4.5. The Occurrence Rate of Giant Planets on Wide
Orbits: Meta-Analysis of Imaging Surveys
The frequency and mass-period distribution of planets
spanning various orbital distances, stellar host masses,
and system ages provides valuable clues about the domi-
nant processes shaping the formation and evolution of
planetary systems. These measurements are best ad-
dressed with large samples and uniform statistical anal-
yses. Nielsen et al. (2008) carried out the first such
large-scale study based on adaptive optics imaging sur-
veys from Biller et al. (2007) and Masciadri et al. (2005).
From their sample of 60 unique stars they found an up-
per limit of 20% for >4 MJup planets between 20–100 AU
at the 95% confidence level. This was expanded to 118
targets in Nielsen & Close (2010) by including the GDPS
survey of Lafreni`ere et al. (2007b), resulting in the same
upper limit and planet mass regime but for a broader
range of separations of 8–911 AU at 68% confidence.
Vigan et al. (2012) and Rameau et al. (2013a) combined
their own observations of high-mass stars with previous
surveys and measured occurrence rates of 8.7+10.1
−2.8 % (for
3–14 MJup planets between 5–320 AU) and 16.1+8.7
−5.3%
(for 1–13 MJup planets between 1–1000 AU), respec-
tively. Brandt et al. (2014a) incorporated the SEEDS,
GDPS, and the NICI moving group surveys and found
a frequency of 1.0–3.1% for 5–70 MJup companions be-
tween 10–100 AU. Recently, Galicher et al. (2016, sub-
mitted) combined results from IDPS, GDPS, and the
NaCo Survey of Young Nearby Austral Stars and found
an occurrence rate of 1.05+2.80
−0.70% for 0.5–14 MJup com-
panions between 20–300 AU based on a sample of 356
unique stars. Breaking this into stellar mass bins did
not reveal any signs of a trend with stellar host mass.
Here I reexamine the occurrence rate of giant planets
with a meta-analysis of the largest and deepest high-
contrast imaging surveys. 696 contrast curves are as-
sembled from the literature from the programs outlined
in Section 4.2. For stars with more than one observa-
tion, the deeper contrast curve at 1′′ is chosen. Targets
with stellar companions within 100 AU are removed from
the sample because binaries can both inhibit planet for-
BA
N=110
FGK
N=155
M
N=118
All
N=384
21
9
.
70
0
.
5
.
0
3
.
0
3
.
0
5
.
0
0.1
0
.
3
0.1
1
.
0
0
.
1
7
.
0
5
.
30
.
0
0.1
100
1
10
100
1
10
100 1000
100
10
1
)
p
u
J
M
(
s
s
a
M
t
e
n
a
P
l
1
10
7
.
0
9
.
0
0.5
0.3
0.1
100
1
10
Semimajor Axis (AU)
n
o
i
t
c
a
r
F
t
e
n
a
P
l
Fig. 11.- Mean sensitivity maps from a meta-analysis of 384 unique stars with published high-contrast imaging observations. M dwarfs
provide the highest sensitivities to lower planet masses in the contrast-limited regime. Altogether, current surveys probe the lowest masses
at separations of ∼30–300 AU. Contours denote 10%, 30%, 50%, 70%, and 90% sensitivity limits.
0.4
0.3
0.2
0.1
0.0
0.0
Direct Imaging
M=5−13 MJup, a=30−300 AU
Radial Velcity
K > 20 m/s, a < 2.5 AU
M
< 3.9%
FGK
< 4.1%
BA
2.8 %+3.7
−2.3
All
0.6 %+0.7
−0.5
0.5
1.0
Stellar Mass (MSun)
1.5
2.0
0.5 1.0 1.5
Fig. 12.- Probability distributions for the occurrence rate giant
planets from a meta-analysis of direct imaging surveys in the liter-
ature. 2.8+3.7
−2.3% of BA stars, <4.1% of FGK stars, and <3.9% of M
dwarfs harbor giant planets between 5–13 MJup and 30–300 AU.
The correlation between stellar host mass and giant planet fre-
quency at small separations (<2.5 AU) from Johnson et al. (2010)
is shown in blue. Larger sample sizes are needed to discern any
such correlation on wide orbits. 0.6+0.7
−0.5% of stars of any mass host
giant planets over the same mass and separation range.
mation and dynamically disturb planetary orbits. Most
candidate planets uncovered during these surveys are re-
jected as background stars from second epoch observa-
tions, but some candidates are either not recovered or
are newly revealed in follow-up data. Because of finite
telescope allocation, some of these candidates remain
untested for common proper motion. These ambigu-
ous candidates cannot be ignored in a statistical anal-
ysis because one (or more) could be indeed be bound.
In these cases, contrast curves are individually trun-
cated one standard deviation above the brightest can-
didate. Ages are taken from the literature except for
members of young moving groups, for which the most
recent (and systematically older) ages of young moving
groups from Bell et al. (2015) are adopted. Most ages
in the sample are less than 300 Myr and within 100 pc.
Altogether this leaves 384 unique stars spanning B2–M6
spectral types: 76 from Bowler et al. (2015b), 72 from
Biller et al. (2013), 61 from Nielsen et al. (2013), 54 from
Lafreni`ere et al. (2007b), 45 from Brandt et al. (2014c),
30 from Janson et al. (2013c), 25 from Vigan et al.
(2012), 14 from Wahhaj et al. (2013b), and 7 from
Janson et al. (2011a).
Sensitivity maps and planet occurrence rates are de-
rived following Bowler et al. (2015b). For a given planet
mass and semimajor axis, a population of artificial plan-
ets on random circular orbits are generated in a Monte
Carlo fashion and converted into apparent magnitudes
and separations using Cond hot-start evolutionary mod-
els from Baraffe et al. (2003), the age of the host star,
and the distance to the system, including uncertainties
in age and distance. These are compared with the mea-
sured contrast curve to infer the fractional sensitivity
at each grid point spanning 30 logarithmically-uniform
bins in mass and separation between 1–1000 AU and
0.5–100 MJup. When available, fractional field of view
coverage is taken into account. Contrasts measured in
CH4S filters are converted to H-band using an empir-
22
104
103
)
h
t
r
a
E
M
(
s
s
a
M
102
10
1
10−2
0.1
VE
1
J
S
U N
Direct Imaging
Radial Velocity
Transit
Microlensing
10
Separation (AU)
102
103
M
a
s
s
(
M
J
u
p
)
10
1
0.1
10−2
10−3
Fig. 13.- The demographics of exoplanets from direct imaging (dark blue), radial velocity (light blue), transit (orange), and microlensing
(green) surveys. Planets detected with radial velocities are minimum masses. It remains unclear whether imaged planets and brown dwarfs
represent distinct populations or whether they form a continuous distribution down to the fragmentation limit. Directly imaged substellar
companions are compiled from the literature, while planets found with other methods are from exoplanets.eu as of April 2016.
ical color-spectral type relationship based on synthetic
colors of ultracool dwarfs from the SpeX Prism Library
(Burgasser 2014) as well as the spectral type-effective
temperature sequence from Golimowski et al. (2004a)5.
The mean sensitivity maps for all 384 targets and sep-
arate bins containing BA stars (110 targets), FGK stars
(155 targets), and M dwarfs (118 targets) are shown in
Figure 11. In general, surveys of high-mass stars probe
higher planet masses than deep imaging around M dwarfs
owing to differences in the host stars' intrinsic luminosi-
ties. The most sensitive region for all stars is between
∼30–300 AU, with less coverage at extremely wide sep-
arations because of limited fields of view and at small
separations in contrast-limited regimes.
The occurrence rate of giant planets for all targets and
for each stellar mass bin are listed in Table 3, which as-
sumes logarithmically-uniform distributions in mass and
separation (see Section 6.5 of Bowler et al. 2015b for de-
tails). The mode and 68.3% minimum credible inter-
val (also known as the highest posterior density inter-
val) of the planet frequency probability distribution are
reported. Two massive stars in the sample host plan-
c2=–0.001542768,
c3=0.0001033761,
5 Synthesized colors of ultracool dwarfs using the Keck/NIRC2
CH4S and HMKO filter profiles yields
the following rela-
tion: CH4S–HMKO = P4
i=0ciSpTi, where c0=0.03913178,
c1=0.008678245,
c4=–
2.902588×10−6, and SpT is the numerical near-infrared spectral
type (M0=1.0, L0=10.0, T0=20.0). This relation is valid from
M3–T8 and the rms of the fit is 0.025 mag. Golimowski et al.
(2004a) provide an empirical Teff (SpT) relationship, but the
inverse SpT(Teff ) is necessary for this filter conversion at a given
mass and age. Refitting the same data from Golimowski et al.
yields the following: SpT = P4
i, where c0=36.56779,
c3=4.108142×10−09,
c1=–0.004666549,
c4=– 4.854263×10−13. This is valid for 700 K < Teff < 3900 K,
the rms is 1.89 mag, and SpT is the same numerical near-infrared
spectral type as above.
i=0ciTeff
c2=–9.872890×10−6,
ets that were either discovered or successfully recovered
in these surveys: β Pic, with a planet at 9 AU, and
HR 8799, with planets spanning 15–70 AU. HR 8799
is treated as a single detection. The most precise oc-
currence rate measurement is between 5–13 MJup and
30–300 AU. Over these ranges, the frequency of plan-
ets orbiting BA, FGK, and M stars is 2.8+3.7
−2.3%, <4.1%,
and <3.9%, respectively (Figure 12). Here upper limits
are 95% confidence intervals. Although there are hints
of a higher giant planet occurrence rate around mas-
sive stars analogous to the well-established correlation at
small separations (Johnson et al. 2007; Lovis & Mayor
2007; Johnson et al. 2010; Bowler et al. 2010b), this
trend is not yet statistically significant at wide orbital
distances and requires larger sample sizes in each stel-
lar mass bin to unambiguously test this correlation.
Marginalizing over stellar host mass, the overall giant
planet occurrence rate for the full sample of 384 stars
is 0.6+0.7
−0.5%, which happens to be comparable to the fre-
quency of hot Jupiters around FGK stars in the field
(1.2 ± 0.4%; Wright et al. 2012) and in the Kepler sam-
ple (0.5 ± 0.1%; Howard et al. 2012). However, com-
pared to the high occurrence rate of giant planets (0.3–
10 MJup) with orbital periods out to 2000 days (∼10%;
Cumming et al. 2008), massive gas giants are clearly
quite rare at wide orbital distances.
5. BROWN DWARFS, GIANT PLANETS, AND THE
COMPANION MASS FUNCTION
Direct imaging has shown that planetary-mass com-
panions exist at unexpectedly wide separations but the
provenance of these objects remains elusive. There is
substantial evidence that the tail-end of the star forma-
tion process can produce objects extending from low-
mass stars at the hydrogen burning limit (≈75 MJup)
to brown dwarfs at the opacity limit for fragmenta-
Indeed,
tion (≈5–10 MJup), which corresponds to the mini-
mum mass of a pressure-supported fragment during the
collapse of a molecular cloud core (Low & Lynden-Bell
1976; Silk 1977; Boss 2001; Bate et al. 2002; Bate
2009).
isolated objects with inferred masses
below 10 MJup have been found in a range of con-
texts over the past decade:
in star-forming regions
(Lucas et al. 2001; Luhman et al. 2009a; Scholz et al.
2012; Muzic et al. 2015), among closer young stel-
lar associations (Liu et al. 2013; Gagn´e et al. 2015a;
Kellogg et al. 2016; Schneider et al. 2016), and at much
older ages as Y dwarfs in the field (Cushing et al. 2011;
Kirkpatrick et al. 2012; Beichman et al. 2013). Simi-
larly, several systems with companions below ≈10 MJup
are difficult to explain with any formation scenario
other than cloud fragmentation: 2M1207–3932 Ab is a
∼25 MJup brown dwarf with a ∼5 MJup companion at
an orbital distance of 40 AU (Chauvin et al. 2004) and
2M0441+2301 AabBab is a quadruple system comprising
a low-mass star, two brown dwarfs, and a 10 MJup object
in a hierarchical and distinctly non-planetary configura-
tion (Todorov et al. 2010).
From the radial velocity perspective, the distribu-
tion of gas giant minimum masses is generally well-fit
with a decaying power law (Butler et al. 2006; Johnson
2009; Lopez & Jenkins 2012) or exponential function
(Jenkins et al. 2016) that tapers off beyond ∼10 MJup.
This is evident in Figure 13, although inhomogeneous ra-
dial velocity detection biases which exclude lower-mass
planets at wide separations are not taken into account.
The dominant formation channel for this population of
close-in giant planets is thought to be core accretion plus
gas capture, in which growing cores reach a critical mass
and undergo runaway gas accretion (e.g., Helled et al.
2013).
The totality of evidence indicates that the decreasing
brown dwarf companion mass function almost certainly
overlaps with the the rising giant planet mass function
in the 5–20 MJup mass range. No strict mass cutoff
can therefore unambiguously divide giant planets from
brown dwarfs, and many of the imaged companions be-
low 13 MJup listed in Table 1 probably originate from
the dwindling brown dwarf companion mass function.
Another approach to separate these populations is to
consider formation channel: planets originate in disks
while brown dwarfs form like stars from the gravita-
tional collapse of molecular cloud cores. However, not
only are the relic signatures of formation difficult to
discern for individual discoveries, but objects spanning
the planetary up to the stellar mass regimes may also
form in large Toomre-unstable circumstellar disks at sep-
arations of tens to hundreds of AU (e.g., Durisen et al.
2007; Kratter & Lodato 2016). Any binary narrative
based on origin in a disk versus a cloud core is there-
fore also problematic. Furthermore, both giant planets
and brown dwarf companions may migrate, dynamically
scatter, or undergo periodic Kozai-Lidov orbital oscil-
lations if a third body is present, further mixing these
populations and complicating the interpretation of very
low-mass companions uncovered with direct imaging.
The deuterium-burning limit at ≈13 MJup is gener-
ally acknowledged as a nebulous, imperfect, and ulti-
mately artificial division between brown dwarfs and gi-
ant planets. Moreover, this boundary is not fixed and
23
may depend on planet composition, core mass, and ac-
cretion history (Spiegel et al. 2011; Bodenheimer et al.
2013; Mordasini 2013). Uncertainties in planet lumi-
nosities, evolutionary histories, metallicities, and ages
can also produce large systematic errors in inferred
planet masses (see Section 3), rendering inconsequen-
tial any sharp boundary set by mass. However, despite
these shortcomings, this border lies in the planet/brown
dwarf "mass valley" and may still serve as a pragmatic
(if flawed) qualitative division between two populations
formed predominantly with their host stars and predom-
inantly in protoplanetary disks.
Observational tests of formation routes will eventually
provide the necessary tools to understand the relation-
ship between these populations. This can be carried out
at an individual level with environmental clues such as
coplanarity of multi-planet systems or orbital alignment
within a debris disk; enhanced metallicities or abundance
ratios relative to host stars (Oberg et al. 2011); or over-
all system orbital architecture. Similarly, the statistical
properties of brown dwarfs and giant planets can be used
to identify dominant formation channels: the separation
distribution of objects formed through cloud fragmenta-
tion should resemble that of binary stars; disk instability
and core accretion may result in a bimodal period distri-
bution for giant planets (Boley 2009); planet scattering
to wide orbits should produce a rising mass function at
low planet masses as opposed to a truncated mass distri-
bution at the fragmentation limit for cloud fragmentation
and disk instability; and the companion mass function
and mass ratio distribution are expected to smoothly ex-
tend from low-mass stars down to the fragmentation limit
if a common formation channel in at play (Brandt et al.
2014a; Reggiani et al. 2016). Testing these scenarios will
require much larger sample sizes given the low occurrence
rates uncovered in direct imaging surveys.
6. CONCLUSIONS AND FUTURE OUTLOOK
High-contrast imaging is still
in its nascence. Ra-
dial velocity, transit, and microlensing surveys have un-
ambiguously demonstrated that giant planets are much
rarer than super-Earths and rocky planets at separations
.10 AU. In that light, the discovery of truly massive
planets at tens, hundreds, and even thousands of AU
with direct imaging is fortuitous, even if the overall oc-
currence rate of this population is quite low. Each detec-
tion technique has produced many micro paradigm shifts
over the past twenty years that disrupt and rearrange
perceptions about the demographics and architectures of
planetary systems. Hot Jupiters, correlations with stellar
mass and metallicity, the ubiquity of super-Earths, com-
pact systems of small planets, resonant configurations,
orbital misalignments, the prevalence of habitable-zone
Earth-sized planets, circumbinary planets, and feature-
less clouds and hazes are an incomplete inventory within
just a few AU (e.g., Winn & Fabrycky 2015). The most
important themes to emerge from direct imaging are that
massive planets exist but are uncommon at wide separa-
tions (>10 AU), and at young ages the low-gravity atmo-
spheres of giant planets do not resemble those of older,
similar-temperature brown dwarfs.
There are many clear directions forward in this
field.
Deeper contrasts and smaller inner working
angles will probe richer portions of planetary mass-
24
and separation distributions. Thirty meter-class tele-
scopes with extreme adaptive optics systems will regu-
larly probe sub-Jovian masses at separations down to
5 AU. This next generation will uncover more plan-
ets and enable a complete mapping of the evolution
of giant planet atmospheres over time. Other fertile
avenues for high-contrast imaging include precise mea-
surements of atmospheric composition (Konopacky et al.
2013; Barman et al. 2015), doppler imaging (Crossfield
2014; Crossfield et al. 2014), photometric monitoring
to map variability of rotationally-modulated features
(e.g., Apai et al. 2016), synergy with other detection
methods (e.g., Lagrange et al. 2013; Sozzetti et al. 2013;
Montet et al. 2014; Clanton & Gaudi 2016), advances in
stellar age-dating at the individual and population lev-
els, merging high-contrast imaging with high-resolution
spectroscopy (Snellen et al. 2014; Snellen et al. 2015),
surveying the companion mass function to sub-Jovian
masses, polarimetric observations of photospheric clouds
(e.g., Marley et al. 2013;
Jensen-Clem et al. 2016),
statistical correlations with stellar host properties,
probing the earliest stages of protoplanet assembly
(Kraus & Ireland 2012; Sallum et al. 2015a), astromet-
ric orbit monitoring and constraints on dynamical his-
tories, and robust dynamical mass measurements to test
evolutionary models and probe initial conditions (e.g.,
Dupuy et al. 2009; Crepp et al. 2012a). High-contrast
imaging has a promising future and will play an ever-
growing role in investigating the architecture, atmo-
spheres, and origin of exoplanets.
It is a pleasure to thank the referee, Rebecca Oppen-
heimer, as well as Lynne Hillenbrand, Dimitri Mawet,
Sasha Hinkley, and Trent Dupuy for their thoughtful
comments and constructive feedback on this review.
Michael Liu, Arthur Vigan, Christian Marois, Motohide
Tamura, Gael Chauvin, Andy Skemer, Adam Kraus, and
Bruce Macintosh contributed helpful suggestions on past
and ongoing imaging surveys. Bruce Macintosh, Eric
Nielsen, Andy Skemer, and Raphael Galicher kindly pro-
vided images for Figure 9. Trent Dupuy generously
shared his compilation of late-T and Y dwarfs used in
Figure 7. This research has made use of the Exoplanet
Orbit Database, the Exoplanet Data Explorer at exo-
planets.org, and the SpeX Prism Spectral Libraries main-
tained by Adam Burgasser. NASA's Astrophysics Data
System Bibliographic Services together with the VizieR
catalogue access tool and SIMBAD database operated at
CDS, Strasbourg, France, were invaluable resources for
this work.
REFERENCES
Absil, O., & Mawet, D. 2009, Astron Astrophys Rev, 18, 317
Absil, O., Le Bouquin, J.-B., Berger, J.-P., et al. 2011, A&A, 535,
A68
Absil, O., Milli, J., Mawet, D., et al. 2013, A&A, 559, L12
Ahmic, M., Jayawardhana, R., Brandeker, A., et al. 2007, The
Astrophysical Journal, 671, 2074
Allard, F., Homeier, D., & Freytag, B. 2012, Philosophical
Transactions of the Royal Society A: Mathematical, Physical
and Engineering Sciences, 370, 2765
Barman, T. S., Macintosh, B., Konopacky, Q. M., & Marois, C.
2011a, The Astrophysical Journal, 733, 65
-. 2011b, The Astrophysical Journal, 735, L39
Basri, G., & Brown, M. E. 2006, Annual Review of Earth and
Planetary Sciences, 34, 193
Bate, M. R. 2009, Monthly Notices RAS, 392, 590
Bate, M. R., Bonnell, I. A., & Bromm, V. 2002, Monthly Notices
RAS, 332, L65
Beichman, C., Gelino, C. R., Kirkpatrick, J. D., et al. 2013, ApJ,
Allers, K. N., & Liu, M. C. 2013, The Astrophysical Journal, 772,
764, 101
79
Allers, K. N., Liu, M. C., Dupuy, T. J., & Cushing, M. C. 2010,
The Astrophysical Journal, 715, 561
Amara, A., & Quanz, S. P. 2012, Monthly Notices RAS, 427, 948
Anderson, E., & Francis, C. 2012, Astronomy Letters, 38, 331
Andrews, S. M. 2015, Publications of the Astronomical Society of
the . . .
-. 2014, ApJ, 783, 68
Bell, C. P. M., Mamajek, E. E., & Naylor, T. 2015, Monthly
Notices RAS, 454, 593
Benedict, G. F., McArthur, B. E., Gatewood, G., et al. 2006, The
Astronomical Journal, 132, 2206
Benest, D., & Duvent, J. L. 1995, A&A, 299, 621
Bergfors, C., Brandner, W., Janson, M., Kohler, R., & Henning,
Andrews, S. M., Rosenfeld, K. A., Kraus, A. L., & Wilner, D. J.
T. 2011, A&A, 528, A134
2013, ApJ, 771, 129
Beust, H., Augereau, J.-C., Bonsor, A., et al. 2014, A&A, 561,
Apai, D., Janson, M., Moro-Mart´ın, A., et al. 2008, The
A43
Astrophysical Journal, 672, 1196
Apai, D., Kasper, M., Skemer, A., et al. 2016, ApJ, 820, 1
Artigau, E., Gagn´e, J., Faherty, J., et al. 2015, ApJ, 806, 1
Aumann, H. H. 1985, Publ. Astron. Soc. Pac., 97, 885
Aumann, H. H., Beichman, C. A., Gillet, F. C., et al. 1984,
Astrophys. J., 278, L23
Beuzit, J.-L., Mouillet, D., Oppenheimer, B. R., & Monnier, J. D.
2007, Protostars and Planets V, 717
Beuzit, J.-L., S´egransan, D., Forveille, T., et al. 2004, A&A, 425,
997
Beuzit, J.-L., Feldt, M., Dohlen, K., et al. 2008, 7014, 701418
Biller, B., Allers, K., Liu, M., Close, L. M., & Dupuy, T. 2011,
Backman, D., Marengo, M., & Stapelfeldt, K. 2009, The
The Astrophysical Journal, 730, 39
Astrophysical . . .
Bailey, J. 2014, Publications of the Astronomical Society of
Australia, 31, e043
Bailey, V., Meshkat, T., Reiter, M., et al. 2014, The Astrophysical
Journal, 780, L4
Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 2002,
A&A, 382, 563
Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., & Hauschildt,
P. H. 2003, A&A, 402, 701
Baraffe, I., Homeier, D., Allard, F., & Chabrier, G. 2015, arXiv,
1503.04107v1
Barenfeld, S. A., Bubar, E. J., Mamajek, E. E., & Young, P. A.
2013, The Astrophysical Journal, 766, 6
Barman, T. S., Konopacky, Q. M., Macintosh, B., & Marois, C.
2015, ApJ, 804, 1
Biller, B., Artigau, ´E., Wahhaj, Z., et al. 2008, Proc. SPIE, 7015,
70156Q
Biller, B., Lacour, S., Juh´asz, A., et al. 2012, The Astrophysical
Journal Letters, 753, L38
Biller, B. A., Close, L. M., Masciadri, E., et al. 2007, The
Astrophysical Journal Supplement Series, 173, 143
Biller, B. A., Liu, M. C., Wahhaj, Z., et al. 2010, The
Astrophysical Journal, 720, L82
-. 2013, ApJ, 777, 160
Biller, B. A., Males, J., Rodigas, T., et al. 2014, The
Astrophysical Journal Letters, 792, L22
Biller, B. A., Vos, J., Bonavita, M., et al. 2015a, The
Astrophysical Journal Letters, 813, 1
Biller, B. A., Liu, M. C., Rice, K., et al. 2015b, Monthly Notices
RAS, 450, 4446
25
Binks, A. S., & Jeffries, R. D. 2014, MNRAS Letters, 438, L11
Binks, A. S., Jeffries, R. D., & Maxted, P. F. L. 2015, Monthly
Caceres, C., Hardy, A., Schreiber, M. R., et al. 2015, The
Astrophysical Journal Letters, 806, 1
Notices RAS, 452, 173
Cantalloube, F., Mouillet, D., Mugnier, L. M., et al. 2015, A&A,
Boccaletti, A., Lagrange, A.-M., Bonnefoy, M., Galicher, R., &
582, A89
Chauvin, G. 2013, A&A, 551, L14
Boccaletti, A., Thalmann, C., Lagrange, A.-M., et al. 2015,
Nature, 526, 230
Bodenheimer, P., D'angelo, G., Lissauer, J. J., Fortney, J. J., &
Saumon, D. 2013, The Astrophysical Journal, 770, 120
Boley, A. C. 2009, ApJL, 695, L53
Boley, A. C., Payne, M. J., Corder, S., et al. 2012, The
Astrophysical Journal, 750, L21
Bond, H. E., Gilliland, R. L., Schaefer, G. H., et al. 2015, ApJ,
813, 1
Cardoso, C. V., Mccaughrean, M. J., King, R. R., et al. 2009, in
AIP Conf. Proc 1094, 15th Cambridge Workshop on CoolStars,
Stellar Systems and the Sun, ed. E. Stempels (Melville, NY:
AIP),, 509
Carson, J., Thalmann, C., Janson, M., et al. 2013, The
Astrophysical Journal, 763, L32
Carson, J. C., Eikenberry, S. S., Brandl, B. R., Wilson, J. C., &
Hayward, T. L. 2005, The Astronomical Journal, 130, 1212
Carson, J. C., Eikenberry, S. S., Smith, J. J., & Cordes, J. M.
2006, The Astronomical Journal, 132, 1146
Bonnefoy, M., Chauvin, G., Rojo, P., et al. 2010, A&A, 512, A52
Bonnefoy, M., Lagrange, A.-M., Boccaletti, A., et al. 2011, A&A,
Chabrier, G. 2001, The Astrophysical Journal, 554, 1274
Chabrier, G., Baraffe, I., Selsis, F., et al. 2007, Protostars and
1
Bonnefoy, M., Boccaletti, A., Lagrange, A.-M., et al. 2013, A&A,
555, A107
Bonnefoy, M., Currie, T., Marleau, G. D., et al. 2014a, A&A, 562,
A111
Planets V, 623
Chabrier, G., Johansen, A., Janson, M., & Rafikov, R. 2014, in
Protostars and Planets VI, Henrik Beuther, Ralf S. Klessen,
Cornelis P. Dullemond, and Thomas Henning (eds.), University
of Arizona Press, Tucson, 914 pp., p.619-642
Bonnefoy, M., Marleau, G. D., Galicher, R., et al. 2014b, A&A,
Chauvin, G., Beust, H., Lagrange, A.-M., & Eggenberger, A.
567, L9
2011, A&A, 528, A8
Bonnefoy, M., Zurlo, A., Baudino, J. L., et al. 2016, A&A, 587,
Chauvin, G., Lagrange, A.-M., Dumas, C., et al. 2004, A&A, 425,
A58
L29
Bonnet-Bidaud, J. M., & Pantin, E. 2008, A&A, 489, 651
Boss, A. P. 2001, The Astrophysical Journal Letters, 551, L167
Bowler, B. P., Andrews, S. M., Kraus, A. L., et al. 2015a, The
Astrophysical Journal Letters, 805, 1
-. 2005a, A&A, 438, L25
Chauvin, G., Lagrange, A.-M., Udry, S., et al. 2006, A&A, 456,
1165
Chauvin, G., Lagrange, A.-M., Udry, S., & Mayor, M. 2007,
Bowler, B. P., & Hillenbrand, L. A. 2015, The Astrophysical
A&A, 475, 723
Journal Letters, 811, L30
Chauvin, G., Thomson, M., Dumas, C., et al. 2003, A&A, 404,
Bowler, B. P., Liu, M. C., Dupuy, T. J., & Cushing, M. C. 2010a,
157
The Astrophysical Journal, 723, 850
Chauvin, G., Lagrange, A.-M., Zuckerman, B., et al. 2005b,
Bowler, B. P., Liu, M. C., Kraus, A. L., & Mann, A. W. 2014,
A&A, 438, L29
ApJ, 784, 65
Chauvin, G., Lagrange, A.-M., Lacombe, F., et al. 2005c, A&A,
Bowler, B. P., Liu, M. C., Kraus, A. L., Mann, A. W., & Ireland,
430, 1027
M. J. 2011, The Astrophysical Journal, 743, 148
Chauvin, G., Lagrange, A.-M., Bonavita, M., et al. 2010, A&A,
Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Dupuy, T. J. 2013,
509, A52
The Astrophysical Journal, 774, 55
Chauvin, G., Lagrange, A.-M., Beust, H., et al. 2012, A&A, 542,
Bowler, B. P., Liu, M. C., Shkolnik, E. L., et al. 2012a, The
A41
Astrophysical Journal, 753, 142
Chauvin, G., Vigan, A., Bonnefoy, M., et al. 2015, A&A, 573,
Bowler, B. P., Liu, M. C., Shkolnik, E. L., & Tamura, M. 2012b,
A127
The Astrophysical Journal, 756, 69
-. 2015b, The Astrophysical Journal Supplement Series, 216, 7
Bowler, B. P., Johnson, J. A., Marcy, G. W., et al. 2010b, The
Astrophysical Journal, 709, 396
Bowler, B. P., Shkolnik, E. L., Liu, M. C., et al. 2015c, The
Astrophysical Journal, 806, 62
Cheetham, A., Hu´elamo, N., Lacour, S., de Gregorio-Monsalvo, I.,
& Tuthill, P. 2015, Monthly Notices RAS Letters, 450, L1
Chiang, E., Kite, E., Kalas, P., Graham, J. R., & Clampin, M.
2009, The Astrophysical Journal, 693, 734
Chun, M., Toomey, D., Wahhaj, Z., et al. 2008, SPIE, 7015, 1
Cieza, L. A., Padgett, D. L., Allen, L. E., et al. 2009, ApJL, 696,
Brandner, W., Zinnecker, H., Alcal´a, J. M., et al. 2000, AJ, 120,
L84
950
Brandt, T. D., McElwain, M. W., Turner, E. L., et al. 2013, The
Astrophysical Journal, 764, 183
-. 2014a, ApJ, 794, 159
Brandt, T. D., McElwain, M. W., Janson, M., et al. 2014b, In
Proc. SPIE, 9148, 1, 91849
Brandt, T. D., Kuzuhara, M., McElwain, M. W., et al. 2014c,
ApJ, 786, 1
Brown, R. A. 2015, ApJ, 805, 1
Bryan, M. L., Knutson, H. A., Howard, A. W., et al. 2016, ApJ,
821, 1
Buenzli, E., Thalmann, C., Vigan, A., et al. 2010, A&A, 524, L1
Burgasser, A. J. 2014, arXiv, 1406.4887v1
Burgasser, A. J., Reid, I. N., Siegler, N., et al. 2007, Protostars
and Planets V, 427
Clanton, C., & Gaudi, B. S. 2016, ApJ, 819, 1
Claudi, R. U., Turatto, M., Gratton, R. G., et al. 2008, in SPIE
Astronomical Telescopes + Instrumentation, ed. I. S. McLean
& M. M. Casali (SPIE), 70143E
Close, L. M., Males, J. R., Kopon, D. A., et al. 2012, Adaptive
Optics Systems III. Proceedings of the SPIE, 8447
Close, L. M., Males, J. R., Morzinski, K., et al. 2013, ApJ, 774, 94
Close, L. M., Follette, K. B., Males, J. R., et al. 2014, The
Astrophysical Journal, 781, L30
Crepp, J. R., Gonzales, E. J., Bechter, E. B., et al. 2016, arXiv,
arXiv:1604.00398
Crepp, J. R., Johnson, J. A., Howard, A. W., et al. 2014, ApJ,
781, 29
-. 2013a, The Astrophysical Journal, 774, 1
Crepp, J. R., Pueyo, L., Brenner, D., et al. 2011, The
Burgasser, A. J., Simcoe, R. A., Bochanski, J. J., et al. 2010, The
Astrophysical Journal, 729, 132
Astrophysical Journal, 725, 1405
Crepp, J. R., Johnson, J. A., Fischer, D. A., et al. 2012a, The
Burningham, B., Leggett, S. K., Homeier, D., et al. 2011,
Astrophysical Journal, 751, 97
Monthly Notices RAS, 414, 3590
Crepp, J. R., Johnson, J. A., Howard, A. W., et al. 2012b, The
Burrows, A., Hubbard, W. B., Lunine, J. I., & Liebert, J. 2001,
Astrophysical Journal, 761, 39
Reviews of Modern Physics, 73, 719
Burrows, A., & Liebert, J. 1993, Reviews of Modern Physics, 65,
301
Burrows, A., Marley, M., Hubbard, W. B., et al. 1997,
Astrophysical Journal, 491, 856
-. 2013b, ApJ, 771, 46
Crepp, J. R., Rice, E. L., Veicht, A., et al. 2015, The
Astrophysical Journal Letters, 798, L43
Crossfield, I. J. M. 2014, A&A, 566, A130
-. 2015, Publications of the Astronomical Society of the Pacific,
Butler, R. P., Wright, J. T., Marcy, G. W., et al. 2006, The
127, 941
Astrophysical Journal, 646, 505
Crossfield, I. J. M., Biller, B., Schlieder, J. E., et al. 2014, Nature,
505, 654
26
Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP,
Ehrenreich, D., Lagrange, A.-M., Montagnier, G., et al. 2010,
120, 531
Currie, T., Bailey, V., Fabrycky, D., et al. 2010, The
Astrophysical Journal, 721, L177
Currie, T., Burrows, A., & Daemgen, S. 2014a, ApJ, 787, 104
Currie, T., Cloutier, R., Brittain, S., et al. 2015, The
Astrophysical Journal Letters, 814, 1
Currie, T., Cloutier, R., Debes, J. H., Kenyon, S. J., & Kaisler,
D. 2013, The Astrophysical Journal Letters, 777, L6
Currie, T., Daemgen, S., Debes, J., et al. 2014b, The
Astrophysical Journal, 780, L30
Currie, T., Fukagawa, M., Thalmann, C., Matsumura, S., &
Plavchan, P. 2012a, The Astrophysical Journal, 755, L34
Currie, T., Thalmann, C., Matsumura, S., et al. 2011a, The
Astrophysical Journal, 736, L33
A&A, 523, A73
Eiroa, C., Marshall, J. P., Mora, A., et al. 2013, A&A, 555, A11
Els, S. G., Sterzik, M. F., Marchis, F., et al. 2001, A&A, 370, L1
Espaillat, C., Muzerolle, J., Najita, J., et al. 2014, Protostars and
Planets VI, 497
Esposito, S., Mesa, D., Skemer, A., et al. 2012, A&A, 549, A52
Fabrycky, D. C., & Murray-Clay, R. A. 2010, The Astrophysical
Journal, 710, 1408
Faherty, J. K., Rice, E. L., Cruz, K. L., Mamajek, E. E., &
N´unez, A. 2012, The Astronomical Journal, 145, 2
Farihi, J., Bond, H. E., Dufour, P., et al. 2013, Monthly Notices
RAS, 430, 652
Feigelson, E. D., Lawson, W. A., & Stark, M. 2006, The
Astronomical . . .
Currie, T., Burrows, A., Itoh, Y., et al. 2011b, The Astrophysical
Fergus, R., Hogg, D. W., Oppenheimer, R., Brenner, D., &
Journal, 729, 128
Currie, T., Debes, J., Rodigas, T. J., et al. 2012b, The
Astrophysical Journal Letters, 760, L32
Currie, T., Burrows, A., Girard, J. H., et al. 2014c, ApJ, 795, 133
Currie, T., Muto, T., Kudo, T., et al. 2014d, The Astrophysical
Journal Letters, 796, L30
Cushing, M. C., Marley, M. S., Saumon, D., et al. 2008, The
Astrophysical Journal, 678, 1372
Pueyo, L. 2014, ApJ, 794, 161
Filippazzo, J. C., Rice, E. L., Faherty, J., et al. 2015, ApJ, 810, 1
Fitzgerald, M. P., Kalas, P. G., & Graham, J. R. 2009, The
Astrophysical Journal, 706, L41
Fortney, J. J., Marley, M. S., Saumon, D., & Lodders, K. 2008,
The Astrophysical Journal, 683, 1104
Fortney, J. J., & Nettelmann, N. 2009, Space Sci Rev, 1
Forveille, T., S gransan, D., Delorme, P., et al. 2004, A&A, 427,
Cushing, M. C., Kirkpatrick, J. D., Gelino, C. R., et al. 2011, The
L1
Astrophysical Journal, 743, 50
Daemgen, S., Bonavita, M., Jayawardhana, R., LaFreniere, D., &
Janson, M. 2015, ApJ, 799, 155
Fuhrmann, K., & Chini, R. 2015, ApJ, 806, 1
Fukagawa, M., Itoh, Y., Tamura, M., et al. 2009, The
Astrophysical Journal, 696, L1
Davies, R., & Kasper, M. 2012, Annu. Rev. Astro. Astrophys., 50,
Gagn´e, J., Burgasser, A. J., Faherty, J. K., et al. 2015a, The
305
Astrophysical Journal Letters, 808, 1
Dawson, R. I., Murray-Clay, R. A., & Fabrycky, D. C. 2011, The
Gagn´e, J., LaFreniere, D., Doyon, R., Malo, L., & Artigau, E.
Astrophysical Journal, 743, L17
De Rosa, R. J., Nielsen, E. L., Blunt, S. C., et al. 2015, The
Astrophysical Journal Letters, 814, 1
De Rosa, R. J., Rameau, J., Patience, J., et al. 2016, arXiv,
1604.01411v1
2014, ApJ, 783, 121
-. 2015b, ApJ, 798, 73
Gagn´e, J., Faherty, J. K., Cruz, K. L., et al. 2015c, The
Astrophysical Journal Supplement Series, 219, 1
Gaidos, E., Mann, A. W., L´epine, S., et al. 2014, Monthly Notices
Deacon, N. R., Schlieder, J. E., & Murphy, S. J. 2016, Monthly
RAS, 443, 2561
Notices RAS, 457, 3191
Galicher, R., Marois, C., Macintosh, B., Barman, T., &
Deacon, N. R., Liu, M. C., Magnier, E. A., et al. 2014, ApJ, 792,
Konopacky, Q. 2011, The Astrophysical Journal, 739, L41
119
Dekany, R., Roberts, J., Burruss, R., et al. 2013, ApJ, 776, 130
Delorme, P., Lagrange, A. M., Chauvin, G., et al. 2012, A&A,
539, A72
Delorme, P., Gagn´e, J., Girard, J. H., et al. 2013, A&A, 553, L5
Dent, W. R. F., Walker, H. J., Holland, W. S., & Greaves, J. S.
2000, Monthly Notices RAS, 314, 702
Dent, W. R. F., Wyatt, M. C., Roberge, A., et al. 2014, Science,
1490
Desidera, S., Covino, E., Messina, S., et al. 2015, A&A, 573, A126
Dohlen, K., Langlois, M., Saisse, M., et al. 2008, in SPIE
Astronomical Telescopes + Instrumentation, ed. I. S. McLean
& M. M. Casali, Lab. d'Astrophysique de Marseille, CNRS,
Univ. de Provence (France) (SPIE), 70143L
Dong, R., Fung, J., & Chiang, E. 2016a, arXiv, 1602.04814v1
Dong, R., Hall, C., Rice, K., & Chiang, E. 2015, The
Astrophysical Journal Letters, 812, L32
Galicher, R., Rameau, J., Bonnefoy, M., et al. 2014, A&A, 565, L4
Garcia, E. V., Dupuy, T. J., Allers, K. N., Liu, M. C., & Deacon,
N. R. 2015, ApJ, 804, 1
Garufi, A., Quanz, S. P., Schmid, H. M., et al. 2016, arXiv,
1601.04983v1
Gauza, B., Bejar, V. J. S., P´erez-Garrido, A., et al. 2015, The
Astrophysical Journal, 804, 96
Geissler, K., Kellner, S., Brandner, W., et al. 2007, A&A, 461, 665
Ginski, C., Mugrauer, M., Seeliger, M., et al. 2016, Monthly
Notices RAS, 457, 2173
Gizis, J. E. 2002, The Astrophysical Journal, 575, 484
Goldman, B., Marsat, S., Henning, T., Clemens, C., & Greiner, J.
2010, Monthly Notices RAS, 405, 1140
Golimowski, D. A., Leggett, S. K., Marley, M. S., et al. 2004a,
The Astronomical Journal, 127, 3516
Golimowski, D. A., Henry, T. J., Krist, J. E., et al. 2004b, The
Astronomical Journal, 128, 1733
Dong, R., Zhu, Z., Fung, J., et al. 2016b, The Astrophysical
Gonzalez, C. A. G., Absil, O., Absil, P. A., et al. 2016, arXiv,
Journal Letters, 816, L12
Dou, J., Ren, D., Zhao, G., et al. 2015, ApJ, 802, 1
Duchene, G. 2010, The Astrophysical Journal, 709, L114
Duchene, G., & Kraus, A. 2013, Annu. Rev. Astro. Astrophys.,
51, 269
1602.08381v1
Go´zdziewski, K., & Migaszewski, C. 2009, Monthly Notices RAS
Letters, 397, L16
-. 2014, Monthly Notices RAS, 440, 3140
Greaves, J. S., Holland, W. S., Moriarty-Schieven, G., et al. 1998,
Dupuy, T. J., Kratter, K. M., Kraus, A. L., et al. 2016, ApJ, 817,
The Astrophysical Journal, 506, L133
1
Dupuy, T. J., & Kraus, A. L. 2013, Science, 341, 1492
Dupuy, T. J., & Liu, M. C. 2011, The Astrophysical Journal, 733,
122
-. 2012, The Astrophysical Journal Supplement, 201, 19
Dupuy, T. J., Liu, M. C., & Ireland, M. J. 2009, The
Astrophysical Journal, 692, 729
-. 2014, ApJ, 790, 133
Dupuy, T. J., Liu, M. C., & Leggett, S. K. 2015a, ApJ, 803, 1
Dupuy, T. J., Liu, M. C., Leggett, S. K., et al. 2015b, ApJ, 805, 1
Durisen, R. H., Boss, A. P., Mayer, L., et al. 2007, in Protostars
and Planets V, e. B. Reipurth, D. Jewitt, & K. Keil (Tucson,
AZ: Univ. Arizona Press), 607
Greaves, J. S., Holland, W. S., Wyatt, M. C., et al. 2005, ApJL,
619, L187
Greco, J. P., & Brandt, T. D. 2016, arXiv, 1602.00691v1
Greco, J. P., & Burrows, A. 2015, ApJ, 808, 1
Guyon, O. 2003, A&A, 404, 379
Guyon, O., Pluzhnik, E. A., Galicher, R., et al. 2005, The
Astrophysical Journal, 622, 744
Guyon, O., Pluzhnik, E. A., Kuchner, M. J., Collins, B., &
Ridgway, S. T. 2006, The Astrophysical Journal Supplement
Series, 167, 81
Hagelberg, J., S´egransan, D., Udry, S., & Wildi, F. 2016, Monthly
Notices RAS, 455, 2178
Han, E., Wang, S. X., Wright, J. T., et al. 2014, Publications of
the Astronomical Society of the Pacific, 126, 827
27
Hashimoto, J., Tamura, M., Muto, T., et al. 2011, The
Janson, M., Brandner, W., Henning, T., et al. 2007, The
Astrophysical Journal, 729, L17
Astronomical Journal, 133, 2442
Hatzes, A. P., Cochran, W. D., McArthur, B., et al. 2000, The
Janson, M., Apai, D., Zechmeister, M., et al. 2009, Monthly
Astrophysical Journal, 544, L145
Notices RAS, 399, 377
Hayward, T. L., Biller, B. A., Liu, M. C., et al. 2014, Publications
of the Astronomical Society of the Pacific, 126, 1112
Heap, S. R., Lindler, D. J., Lanz, T. M., et al. 2000, The
Janson, M., Carson, J., Thalmann, C., et al. 2011b, ApJ, 728, 85
Janson, M., Brandt, T. D., Kuzuhara, M., et al. 2013b, The
Astrophysical Journal, 778, L4
Astrophysical Journal, 539, 435
Janson, M., Brandt, T. D., Moro-Mart´ın, A., et al. 2013c, ApJ,
Heinze, A. N., Hinz, P. M., Kenworthy, M., et al. 2010a, The
773, 73
Astrophysical Journal, 714, 1570
Heinze, A. N., Hinz, P. M., Kenworthy, M., Miller, D., &
Sivanandam, S. 2008, The Astrophysical Journal, 688, 583
Heinze, A. N., Hinz, P. M., Sivanandam, S., et al. 2010b, The
Astrophysical Journal, 714, 1551
Janson, M., Bergfors, C., Brandner, W., et al. 2014, The
Astrophysical Journal Supplement Series, 214, 17
Jeffries, R. D. 2014, EAS Publications Series, 65, 289
Jenkins, J. S., Jones, H. R. A., Biller, B., et al. 2010, A&A, 515,
A17
Helled, R., Bodenheimer, P., Podolak, M., et al. 2013, in
Jenkins, J. S., Jones, H. R. A., Tuomi, M., et al. 2016, arXiv,
Protostars and Planets VI, ed. H. Beuther et al. (Tucson, AZ:
Univ. Arizona Press), 643
Helling, C., & Casewell, S. 2014, Astron Astrophys Rev, 22, 80
Helling, C., Woitke, P., Rimmer, P., et al. 2014, Life, 4, 142
Helling, C., Ackerman, A., Allard, F., et al. 2008, Monthly
Notices RAS, 391, 1854
1603.09391v1
Jensen-Clem, R., Millar-Blanchaer, M., Mawet, D., et al. 2016,
ApJ, 820, 1
Jilkova, L., & Zwart, S. P. 2015, Monthly Notices RAS, 451, 804
Johnson, J. A. 2009, Publications of the Astronomical Society of
the Pacific, 121, 309
Henry, T. J., & McCarthy, D. W. J. 1990, Astrophysical Journal,
Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R.
350, 334
Herczeg, G. J., & Hillenbrand, L. A. 2015, The Astrophysical
Journal, 808, 23
2010, Publications of the Astronomical Society of the Pacific,
122, 905
Johnson, J. A., Butler, R. P., Marcy, G. W., et al. 2007, The
Hinkley, S., Oppenheimer, B. R., Soummer, R., et al. 2007, The
Astrophysical Journal, 670, 833
Astrophysical Journal, 654, 633
Jones, J., White, R. J., Quinn, S., et al. 2016, The Astrophysical
Hinkley, S., Oppenheimer, B. R., Brenner, D., et al. 2010, The
Journal Letters, 822, 1
Astrophysical Journal, 712, 421
Jovanovic, N., Martinache, F., Guyon, O., et al. 2015,
Hinkley, S., Oppenheimer, B. R., Zimmerman, N., et al. 2011,
Publications of the Astronomical Society of Pacific, 127, 890
Publications of the Astronomical Society of the Pacific, 123, 74
Kalas, P., Graham, J., Chiang, E., & Fitzgerald, M. 2008a,
Hinkley, S., Pueyo, L., Faherty, J. K., et al. 2013, ApJ, 779, 153
Hinkley, S., Bowler, B. P., Vigan, A., et al. 2015a, The
Astrophysical Journal Letters, 805, 1
Science Express
Kalas, P., Graham, J. R., & Clampin, M. 2005, Nature, 435, 1067
Kalas, P., Graham, J. R., Fitzgerald, M. P., & Clampin, M. 2013,
Hinkley, S., Kraus, A. L., Ireland, M. J., et al. 2015b, The
ApJ, 775, 56
Astrophysical Journal Letters, 806, 1
Kalas, P., & Jewitt, D. 1995, Astronomical Journal v.110, 110,
Hinz, P. M., Heinze, A. N., Sivanandam, S., et al. 2006, ApJ, 653,
794
1486
Kalas, P., Graham, J. R., Chiang, E., et al. 2008b, Science, 322,
Hinz, P. M., Rodigas, T. J., Kenworthy, M. A., et al. 2010, The
1345
Astrophysical Journal, 716, 417
Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, The
Astrophysical Journal, 721, 1467
Kalas, P. G., Rajan, A., Wang, J. J., et al. 2015, ApJ, 814, 1
Kasper, M., Apai, D., Janson, M., & Brandner, W. 2007, A&A,
472, 321
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, The
Kellogg, K., Metchev, S., Gagn´e, J., & Faherty, J. 2016, The
Astrophysical Journal Supplement Series, 201, 15
Astrophysical Journal Letters, 821, 1
Howard, A. W., Marcy, G. W., Fischer, D. A., et al. 2014, ApJ,
Kennedy, G. M., & Wyatt, M. C. 2011, Monthly Notices RAS,
794, 51
Huby, E., Perrin, G., Marchis, F., et al. 2012, A&A, 541, A55
Hu´elamo, N., Lacour, S., Tuthill, P., et al. 2011, A&A, 528, L7
Hung, L.-W., Duchene, G., Arriaga, P., et al. 2015, The
Astrophysical Journal Letters, 815, 1
412, 2137
Kenworthy, M. A., Codona, J. L., Hinz, P. M., et al. 2007, ApJ,
660, 762
Kenworthy, M. A., Mamajek, E. E., Hinz, P. M., et al. 2009, The
Astrophysical Journal, 697, 1928
Ingraham, P., Marley, M. S., Saumon, D., et al. 2014, The
Kenworthy, M. A., Meshkat, T., Quanz, S. P., et al. 2013, ApJ,
Astrophysical Journal Letters, 794, L15
764, 7
Ireland, M. J., Kraus, A., Martinache, F., Law, N., &
Hillenbrand, L. A. 2011, The Astrophysical Journal, 726, 113
Ireland, M. J., & Kraus, A. L. 2008, The Astrophysical Journal,
Kenyon, S. J., & Bromley, B. C. 2004, The Astronomical Journal
Kenyon, S. J., Currie, T., & Bromley, B. C. 2014, ApJ, 786, 70
Kirkpatrick, J. D., Gelino, C. R., Cushing, M. C., et al. 2012, The
678, L59
-. 2014, IAU, 8, 199
Irwin, A. W., Fletcher, J. M., Yang, S. L. S., Walker, G. A. H., &
Goodenough, C. 1992, Astronomical Society of the Pacific, 104,
489
Isella, A., Chandler, C. J., Carpenter, J. M., P´erez, L. M., &
Ricci, L. 2014, ApJ, 788, 129
Astrophysical Journal, 753, 156
Kiss, L. L., Mo´or, A., Szalai, T., et al. 2010, Monthly Notices
RAS, 411, 117
Knutson, H. A., Fulton, B. J., Montet, B. T., et al. 2014, ApJ,
785, 126
Konopacky, Q. M., Barman, T. S., Macintosh, B. A., & Marois,
C. 2013, Science, 339, 1398
Itoh, Y., Oasa, Y., & Fukagawa, M. 2006, The Astrophysical
Konopacky, Q. M., Marois, C., Macintosh, B. A., et al. 2016,
Journal
Itoh, Y., Hayashi, M., Tamura, M., et al. 2005, The Astrophysical
Journal, 620, 984
arXiv, 1604.08157v1
Kratter, K. M., & Lodato, G. 2016, arXiv, 1603.01280v1
Kraus, A. L., Andrews, S. M., Bowler, B. P., et al. 2015, The
Itoh, Y., Tamura, M., Hayashi, M., et al. 2008, Publications of
Astrophysical Journal, 798, L23
the Astronomical Society of Japan, 60, 209
Janson, M., Bonavita, M., Klahr, H., et al. 2011a, The
Astrophysical Journal, 736, 89
Kraus, A. L., & Hillenbrand, L. A. 2012, The Astrophysical
Journal, 757, 141
Kraus, A. L., & Ireland, M. J. 2012, The Astrophysical Journal,
Janson, M., Carson, J. C., Lafreni`ere, D., et al. 2012, The
745, 5
Astrophysical Journal, 747, 116
Kraus, A. L., Ireland, M. J., Cieza, L. A., et al. 2014a, ApJ, 781,
Janson, M., LaFreniere, D., Jayawardhana, R., et al. 2013a, ApJ,
20
773, 170
Kraus, A. L., Ireland, M. J., Hillenbrand, L. A., & Martinache, F.
Janson, M., Quanz, S. P., Carson, J. C., et al. 2015, A&A, 574,
2012, The Astrophysical Journal, 745, 19
A120
Kraus, A. L., Ireland, M. J., Huber, D., Mann, A. W., & Dupuy,
Janson, M., Reffert, S., Brandner, W., et al. 2008, A&A, 488, 771
T. J. 2016, arXiv, 1604.05744v1
28
Kraus, A. L., Ireland, M. J., Martinache, F., & Lloyd, J. P. 2008,
Lowrance, P. J., Schneider, G., Kirkpatrick, J. D., et al. 2000,
The Astrophysical Journal, 679, 762
The Astrophysical Journal, 541, 390
Kraus, A. L., Shkolnik, E. L., Allers, K. N., & Liu, M. C. 2014b,
Lowrance, P. J., Becklin, E. E., Schneider, G., et al. 2005, The
AJ, 147, 146
Astronomical Journal, 130, 1845
Kraus, A. L., White, R. J., & Hillenbrand, L. A. 2005, ApJ, 633,
Lucas, P. W., Roche, P. F., Allard, F., & Hauschildt, P. H. 2001,
452
Krivov, A. V. 2010, Research in Astron. Astrophys., 383
Kuchner, M. J., & Brown, M. E. 2000, Publications of the
Astronomical Society of the Pacific, 112, 827
Kumar, S. S. 1963, Astrophysical Journal, 137, 1121
Kuzuhara, M., Tamura, M., Ishii, M., et al. 2011, The
Astronomical Journal, 141, 119
Monthly Notices RAS, 326, 695
Luhman, K. L., Burgasser, A. J., & Bochanski, J. J. 2011, The
Astrophysical Journal, 730, L9
Luhman, K. L., Burgasser, A. J., Labb´e, I., et al. 2012, The
Astrophysical Journal, 744, 135
Luhman, K. L., & Jayawardhana, R. 2002, The Astrophysical
Journal, 566, 1132
Kuzuhara, M., Tamura, M., Kudo, T., et al. 2013, The
Luhman, K. L., Mamajek, E. E., Allen, P. R., & Cruz, K. L.
Astrophysical Journal, 774, 11
Lachapelle, F.-R., LaFreniere, D., Gagn´e, J., et al. 2015, ApJ,
802, 1
Lacour, S., Biller, B., Cheetham, A., et al. 2015, arXiv,
1511.09390v1
2009a, The Astrophysical Journal, 703, 399
Luhman, K. L., Mamajek, E. E., Allen, P. R., Muench, A. A., &
Finkbeiner, D. P. 2009b, The Astrophysical Journal, 691, 1265
Luhman, K. L., Mcleod, K. K., & Goldenson, N. 2005, The
Astrophysical Journal, 623, 1141
Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk, M. H. 2008,
Luhman, K. L., Wilson, J. C., Brandner, W., et al. 2006, The
The Astrophysical Journal, 689, L153
-. 2010, The Astrophysical Journal, 719, 497
Lafreni`ere, D., Jayawardhana, R., van Kerkwijk, M. H.,
Brandeker, A., & Janson, M. 2014, ApJ, 785, 47
Lafreni`ere, D., Marois, C., Doyon, R., & Barman, T. 2009, The
Astrophysical Journal, 694, L148
Lafreni`ere, D., Marois, C., Doyon, R., Nadeau, D., & Artigau, ´E.
2007a, The Astrophysical Journal, 660, 770
Lafreni`ere, D., Doyon, R., Marois, C., et al. 2007b, The
Astrophysical Journal, 670, 1367
Lagrange, A.-M. 2014, Philosophical Transactions of the Royal
Society A: Mathematical, Physical and Engineering Sciences,
372, 20130090
Lagrange, A.-M., Beust, H., Udry, S., Chauvin, G., & Mayor, M.
2006, A&A, 459, 955
Lagrange, A.-M., Meunier, N., Chauvin, G., et al. 2013, A&A,
559, A83
Lagrange, A.-M., Gratadour, D., Chauvin, G., et al. 2009a, A&A,
493, L21
Lagrange, A.-M., Kasper, M., Boccaletti, A., et al. 2009b, A&A,
506, 927
Lagrange, A.-M., Bonnefoy, M., Chauvin, G., et al. 2010, Science,
329, 57
Lagrange, A.-M., Langlois, M., Gratton, R., et al. 2016, A&A,
586, L8
Langlois, M., Dohlen, K., Vigan, A., et al. 2014, in SPIE
Astronomical Telescopes + Instrumentation, ed. S. K. Ramsay,
I. S. McLean, & H. Takami (SPIE), 91471R
Larwood, J. D., & Kalas, P. G. 2001, MNRAS, 323, 402
Leconte, J., Soummer, R., Hinkley, S., et al. 2010, The
Astrophysical Journal, 716, 1551
Lee, J.-M., Heng, K., & Irwin, P. G. J. 2013, ApJ, 778, 97
L´epine, S., & Simon, M. 2009, The Astronomical Journal, 137,
3632
Line, M. R., Fortney, J. J., Marley, M. S., & Sorahana, S. 2014,
ApJ, 793, 33
Liu, M. C. 2004, Science, 305, 1442
Liu, M. C., Dupuy, T. J., Bowler, B. P., Leggett, S. K., & Best,
W. M. J. 2012, The Astrophysical Journal, 758, 57
Liu, M. C., Fischer, D. A., Graham, J. R., et al. 2002, The
Astrophysical Journal, 571, 519
Liu, M. C., Wahhaj, Z., Biller, B. A., et al. 2010, SPIE, 7736,
77361K
Liu, M. C., Delorme, P., Dupuy, T. J., et al. 2011, The
Astrophysical Journal, 740, 108
Liu, M. C., Magnier, E. A., Deacon, N. R., et al. 2013, The
Astrophysical Journal, 777, L20
Lodato, G., Delgado-Donate, E., & Clarke, C. J. 2005, Monthly
Notices RAS Letters, 364, L91
Looper, D. L., Mohanty, S., Bochanski, J. J., et al. 2010, The
Astrophysical Journal, 714, 45
Lopez, S., & Jenkins, J. S. 2012, ApJ, 756, 177
Lovis, C., & Mayor, M. 2007, A&A, 472, 657
Low, C., & Lynden-Bell, D. 1976, Royal Astronomical Society,
176, 367
Lowrance, P. J., McCarthy, C., Becklin, E. E., et al. 1999, The
Astrophysical Journal, 512, L69
Astrophysical Journal, 649, 894
Luhman, K. L., Patten, B. M., Marengo, M., et al. 2007, The
Astrophysical Journal, 654, 570
Macintosh, B., Graham, J. R., Ingraham, P., et al. 2014,
Proceedings of the National Academy of Sciences, 111, 12661
Macintosh, B., Graham, J. R., Barman, T., et al. 2015, Science
Macintosh, B. A., Becklin, E. E., Kaisler, D., Konopacky, Q., &
Zuckerman, B. 2003, The Astrophysical Journal, 594, 538
Macintosh, B. A., Max, C., Zuckerman, B., et al. 2001, Young
Stars Near Earth: Progress and Prospects, 244, 309
Madhusudhan, N., Knutson, H., Fortney, J., & Barman, T. 2014,
arXiv, 1402.1169v1
Maire, A. L., Boccaletti, A., Rameau, J., et al. 2014, A&A, 566,
A126
Maire, A. L., Skemer, A. J., Hinz, P. M., et al. 2015a, A&A, 576,
A133
-. 2015b, A&A, 579, C2
Maire, A. L., Bonnefoy, M., Ginski, C., et al. 2016, A&A, 587,
A56
Males, J. R., Close, L. M., Morzinski, K. M., et al. 2014, ApJ,
786, 32
Malo, L., Artigau, E., Doyon, R., et al. 2014, ApJ, 788, 81
Malo, L., Doyon, R., Lafreni`ere, D., et al. 2013, ApJ, 762, 88
Mamajek, E. E. 2012, The Astrophysical Journal, 754, L20
-. 2016, In Young Stars & Planets Near the Sun, Proceedings of
the International Astronomical Union, IAU Symposium, 314, 21
Mamajek, E. E., & Bell, C. P. M. 2014, Monthly Notices RAS,
445, 2169
Mamajek, E. E., Bartlett, J. L., Seifahrt, A., et al. 2013, The
Astronomical Journal, 146, 154
Marengo, M., Megeath, S. T., Fazio, G. G., et al. 2006, The
Astrophysical Journal, 647, 1437
Marengo, M., Stapelfeldt, K., Werner, M. W., et al. 2009, The
Astrophysical Journal, 700, 1647
Marleau, G. D., & Cumming, A. 2013, Monthly Notices RAS,
437, 1378
Marley, M. S., Ackerman, A. S., Cuzzi, J. N., & Kitzmann, D.
2013, arXiv, astro-ph.EP
Marley, M. S., Fortney, J., Seager, S., & Barman, T. 2007a,
Protostars and Planets V, 733
Marley, M. S., Fortney, J. J., Hubickyj, O., Bodenheimer, P., &
Lissauer, J. J. 2007b, The Astrophysical Journal, 655, 541
Marley, M. S., & Robinson, T. D. 2015, Annu. Rev. Astro.
Astrophys., 53, 279
Marley, M. S., Saumon, D., Cushing, M., et al. 2012, The
Astrophysical Journal, 754, 135
Marois, C., Correia, C., Galicher, R., et al. 2014, in SPIE
Astronomical Telescopes + Instrumentation, ed. E. Marchetti,
L. M. Close, & J.-P. Veran (SPIE), 91480U
Marois, C., Lafreni`ere, D., Doyon, R., Macintosh, B., & Nadeau,
D. 2006, The Astrophysical Journal, 641, 556
Marois, C., Macintosh, B., Barman, T., et al. 2008, Science, 322,
1348
Marois, C., Macintosh, B., & V´eran, J.-P. 2010a, Proc. SPIE,
7736, 77361J
Marois, C., Zuckerman, B., Konopacky, Q. M., Macintosh, B., &
Barman, T. 2010b, Nature, 468, 1080
29
Marshall, J. P., Moro-Martin, A., Eiroa, C., et al. 2014, A&A,
Mugrauer, M., & Neuhauser, R. 2005, Monthly Notices RAS
565, A15
Letters, 361, L15
Martinache, F., Guyon, O., Jovanovic, N., et al. 2014,
Mugrauer, M., Neuh User, R., Guenther, E. W., et al. 2004,
Publications of the Astronomical Society of Pacific, 126, 565
A&A, 417, 1031
Masciadri, E., Mundt, R., Henning, T., Alvarez, C., & barrado y
Muto, T., Grady, C. A., Hashimoto, J., et al. 2012, The
Navascu´es, D. 2005, The Astrophysical Journal, 625, 1004
Astrophysical Journal, 748, L22
Matthews, B. C., Krivov, A. V., Wyatt, M. C., Bryden, G., &
Muzic, K., Scholz, A., Geers, V. C., & Jayawardhana, R. 2015,
Eiroa, C. 2014a, in Protostars and Planets VI, Henrik Beuther,
Ralf S. Klessen, Cornelis P. Dullemond, and Thomas Henning
(eds.), University of Arizona Press, Tucson, 521
ApJ, 810, 1
Nakajima, T., Durrance, S. T., Golimowski, D. A., & Kulkarni,
S. R. 1994, The Astrophysical Journal, 428, 797
Matthews, C. T., Crepp, J. R., Skemer, A., et al. 2014b, The
Nakajima, T., Morino, J.-I., Tsuji, T., et al. 2005, Astronomische
Astrophysical Journal Letters, 783, L25
Mawet, D., Riaud, P., Absil, O., & Surdej, J. 2005, The
Astrophysical Journal
Nachrichten, 326, 952
Naud, M.-E., Artigau, E., Malo, L., et al. 2014, ApJ, 787, 5
Neuhauser, R., Guenther, E. W., Alves, J., et al. 2003,
Mawet, D., Serabyn, E., Liewer, K., et al. 2010, The
Astronomische Nachrichten, 324, 535
Astrophysical Journal, 709, 53
Neuhauser, R., Mugrauer, M., Fukagawa, M., Torres, G., &
Mawet, D., Absil, O., Montagnier, G., et al. 2012a, A&A, 544,
Schmidt, T. 2007, A&A, 462, 777
A131
Mawet, D., Pueyo, L., Lawson, P., et al. 2012b, in SPIE
Astronomical Telescopes + Instrumentation, ed. M. C.
Clampin, G. G. Fazio, H. A. MacEwen, & J. M. Oschmann
(SPIE), 844204
Mawet, D., Milli, J., Wahhaj, Z., et al. 2014, ApJ, 792, 97
Mawet, D., David, T., Bottom, M., et al. 2015, ApJ, 811, 1
Mayama, S., Hashimoto, J., Muto, T., et al. 2012, The
Astrophysical Journal Letters, 760, L26
McCarthy, C., & Zuckerman, B. 2004, The Astronomical Journal,
127, 2871
Ngo, H., Knutson, H. A., Hinkley, S., et al. 2015, ApJ, 800, 138
Nielsen, E. L., & Close, L. M. 2010, The Astrophysical Journal,
717, 878
Nielsen, E. L., Close, L. M., Biller, B. A., Masciadri, E., &
Lenzen, R. 2008, The Astrophysical Journal, 674, 466
Nielsen, E. L., Liu, M. C., Wahhaj, Z., et al. 2012, The
Astrophysical Journal, 750, 53
-. 2013, ApJ, 776, 4
-. 2014, ApJ, 794, 158
Norris, B., Schworer, G., Tuthill, P., et al. 2015, Monthly Notices
RAS, 447, 2894
Mccarthy, K., & Wilhelm, R. J. 2014, The Astronomical Journal,
Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011, The
148, 70
Mennesson, B., Serabyn, E., Hanot, C., et al. 2011, ApJ, 736, 14
Meshkat, T., Bailey, V. P., Su, K. Y. L., et al. 2015a, ApJ, 800, 5
Meshkat, T., Kenworthy, M. A., Quanz, S. P., & Amara, A.
2013a, ApJ, 780, 17
Astrophysical Journal, 743, L16
Olofsson, J., Benisty, M., Le Bouquin, J.-B., et al. 2013, A&A,
552, A4
Olofsson, J., Samland, M., Avenhaus, H., et al. 2016, arXiv,
1601.07861v1
Meshkat, T., Kenworthy, M. A., Reggiani, M., et al. 2015b,
Oppenheimer, B. R., Golimowski, D. A., Kulkarni, S. R., et al.
Monthly Notices RAS, 453, 2534
2001, The Astronomical Journal, 121, 2189
Meshkat, T., Bailey, V., Rameau, J., et al. 2013b, The
Oppenheimer, B. R., & Hinkley, S. 2009, Annu. Rev. Astro.
Astrophysical Journal, 775, L40
Meshkat, T., Bonnefoy, M., Mamajek, E. E., et al. 2015c,
Monthly Notices RAS, 453, 2379
Metchev, S., Marois, C., & Zuckerman, B. 2009, The
Astrophysical Journal, 705, L204
Metchev, S. A., & Hillenbrand, L. A. 2004, The Astrophysical
Journal, 617, 1330
-. 2006, The Astrophysical Journal, 651, 1166
-. 2009, The Astrophysical Journal Supplement Series, 181, 62
Metchev, S. A., Hillenbrand, L. A., & White, R. J. 2003, The
Astrophysical Journal, 582, 1102
Metchev, S. A., Heinze, A., Apai, D., et al. 2015, ApJ, 799, 154
Meyer, M. R., Carpenter, J. M., Mamajek, E. E., et al. 2008,
ApJL, 673, L181
Astrophys., 47, 253
Oppenheimer, B. R., Kulkarni, S. R., & Stauffer, J. R. 2000,
Protostars and Planets IV (Book - Tucson: University of
Arizona Press; eds Mannings, 1313
Oppenheimer, B. R., Digby, A. P., Newburgh, L., et al. 2004, in
Astronomical Telescopes and Instrumentation, ed.
D. Bonaccini Calia, B. L. Ellerbroek, & R. Ragazzoni (SPIE),
433–442
Oppenheimer, B. R., Baranec, C., Beichman, C., et al. 2013, The
Astrophysical Journal, 768, 24
Owen, J. E. 2016, Publications of the Astronomical Society of
Australia, 33, e005
Owen, J. E., & Menou, K. 2016, The Astrophysical Journal
Letters, 819, 1
Millar-Blanchaer, M. A., Graham, J. R., Pueyo, L., et al. 2015,
Ozernoy, L. M., Gorkavyi, N. N., Mather, J. C., & Taidakova,
ApJ, 811, 1
Milli, J., Lagrange, A.-M., Mawet, D., et al. 2014, A&A, 566, A91
Mohanty, S., Greaves, J., Mortlock, D., et al. 2013, ApJ, 773, 168
Molli`ere, P., & Mordasini, C. 2012, A&A, 547, A105
Montet, B. T., Crepp, J. R., Johnson, J. A., Howard, A. W., &
Marcy, G. W. 2014, ApJ, 781, 28
T. A. 2000, The Astrophysical Journal, 537, L147
Patience, J., King, R. R., De Rosa, R. J., & Marois, C. 2010,
A&A, 517, A76
Pecaut, M. J., Mamajek, E. E., & Bubar, E. J. 2012, The
Astrophysical Journal, 746, 154
Perez, S., Dunhill, A., Casassus, S., et al. 2015, The Astrophysical
Montet, B. T., Bowler, B. P., Shkolnik, E. L., et al. 2015, The
Journal Letters, 811, 1
Astrophysical Journal Letters, 813, 1
Perryman, M. 2011, The Exoplanet Handbook (Cambridge
Moor, A., Szabo, G. M., Kiss, L. L., et al. 2013, Monthly Notices
University Press)
RAS, 1
Mordasini, C. 2013, A&A, 558, A113
Mordasini, C., Alibert, Y., Klahr, H., & Henning, T. 2012, A&A,
547, A111
Moro-Martin, A., Wyatt, M. C., Malhotra, R., & Trilling, D. E.
2008, The Solar System Beyond Neptune, 465
Moro-Martin, A., Marshall, J. P., Kennedy, G., et al. 2015, ApJ,
801, 1
Morzinski, K. M., Close, L. M., Males, J. R., et al. 2014, in SPIE
Astronomical Telescopes + Instrumentation, ed. E. Marchetti,
L. M. Close, & J.-P. Veran (SPIE), 914804
Morzinski, K. M., Males, J. R., Skemer, A. J., et al. 2015, ApJ,
815, 1
Mouillet, D., Larwood, J. D., Papaloizou, J. C. B., & Lagrange,
A.-M. 1997, Monthly Notices RAS, 292, 896
Perryman, M., Hartman, J., Bakos, G. ´A., & Lindegren, L. 2014,
ApJ, 797, 14
Petr-Gotzens, M. G., Cuby, J. G., Smith, M. D., & Sterzik, M. F.
2010, A&A, 522, A78
Pueyo, L., Crepp, J. R., Vasisht, G., et al. 2012, The
Astrophysical Journal Supplement Series, 199, 6
Pueyo, L., Soummer, R., Hoffmann, J., et al. 2015, ApJ, 803, 31
Quanz, S. P. 2015, Astrophys. Space Sci., 1
Quanz, S. P., Amara, A., Meyer, M. R., et al. 2015, ApJ, 807, 1
-. 2013, The Astrophysical Journal, 766, L1
Quanz, S. P., Meyer, M. R., Kenworthy, M. A., et al. 2010, The
Astrophysical Journal Letters, 722, L49
Queloz, D., Mayor, M., Weber, L., et al. 2000, A&A, 354, 99
Radigan, J., LaFreniere, D., Jayawardhana, R., & Artigau, E.
2014, ApJ, 793, 75
30
Rajan, A., Barman, T., Soummer, R., et al. 2015, The
Shkolnik, E. L., Anglada-Escud´e, G., Liu, M. C., et al. 2012, The
Astrophysical Journal Letters, 809, 1
Rameau, J., Chauvin, G., Lagrange, A.-M., et al. 2015, A&A
-. 2013a, A&A, 553, A60
-. 2013b, The Astrophysical Journal, 779, L26
-. 2013c, The Astrophysical . . . , 772, L15
Rameau, J., Nielsen, E. L., De Rosa, R. J., et al. 2016, arXiv,
1604.05139v1
Rebolo, R., Zapatero Osorio, M. R., Madruga, S., et al. 1998,
Science, 282, 1309
Reggiani, M., Quanz, S. P., Meyer, M. R., et al. 2014, The
Astrophysical Journal Letters, 792, L23
Astrophysical Journal, 758, 56
Silk, J. 1977, Astrophysical Journal, 214, 152
Skemer, A. J., & Close, L. M. 2011, ApJ, 730, 53
Skemer, A. J., Marley, M. S., Hinz, P. M., et al. 2014a, ApJ, 792,
17
Skemer, A. J., Hinz, P., Esposito, S., et al. 2014b, in SPIE
Astronomical Telescopes + Instrumentation, ed. E. Marchetti,
L. M. Close, & J.-P. Veran (SPIE), 91480L
Skemer, A. J., Hinz, P., Montoya, M., et al. 2015, in SPIE Optical
Engineering + Applications, ed. S. Shaklan (SPIE), 96051D
Skemer, A. J., Morley, C. V., Zimmerman, N. T., et al. 2016,
Reggiani, M., Meyer, M. R., Chauvin, G., et al. 2016, A&A, 586,
ApJ, 817, 1
A147
Reid, I. N., Cruz, K. L., Kirkpatrick, J. D., et al. 2008, The
Astronomical Journal, 136, 1290
Reid, I. N., & Walkowicz, L. M. 2006, The Publications of the
Astronomical Society of the Pacific, 118, 671
Riaz, B., & Mart´ın, E. L. 2011, A&A, 525, A10
Riaz, B., Mart´ın, E. L., Petr-Gotzens, M. G., & Monin, J.-L.
2013, A&A, 559, A109
Rice, E. L., Oppenheimer, R., Zimmerman, N., Roberts, Jr, L. C.,
& Hinkley, S. 2015, Publications of the Astronomical Society of
the Pacific, 127, 479
Riedel, A. R., Finch, C. T., Henry, T. J., et al. 2014, The
Astronomical Journal, 147, 85
Smith, B. A., & Terrile, R. J. 1984, Science
Snellen, I., de Kok, R., Birkby, J. L., et al. 2015, A&A, 576, A59
Snellen, I. A. G., Brandl, B. R., de Kok, R. J., et al. 2014,
Nature, 508, 63
Soderblom, D. R. 2010, Annu. Rev. Astro. Astrophys., 48, 581
Soderblom, D. R., Hillenbrand, L. A., Jeffries, R. D., Mamajek,
E. E., & Naylor, T. 2014, . . . and Planets VI
Song, I., Zuckerman, B., & Bessell, M. S. 2003, The Astrophysical
Journal, 599, 342
Soummer, R. 2005, The Astrophysical Journal, 618, L161
Soummer, R., Hagan, J. B., Pueyo, L., et al. 2011, The
Astrophysical Journal, 741, 55
Soummer, R., Pueyo, L., & Larkin, J. 2012, The Astrophysical
Rieke, G. H., Su, K. Y. L., Stansberry, J. A., et al. 2005, The
Journal, 755, L28
Astrophysical Journal, 620, 1010
Sozzetti, A., Giacobbe, P., Lattanzi, M. G., et al. 2013, Monthly
Roberts, L. C., Rice, E. L., Beichman, C. A., et al. 2012, The
Notices RAS, 437, 497
Astronomical Journal, 144, 14
Robinson, T. D., Stapelfeldt, K. R., & Marley, M. S. 2016,
Spergel, D., Gehrels, N., Baltay, C., et al. 2015, arXiv, 1
Spiegel, D. S., & Burrows, A. 2012, The Astrophysical Journal,
Publications of the Astronomical Society of the Pacific, 128, 1
745, 174
Rodigas, T. J., Follette, K. B., Weinberger, A., Close, L., &
Spiegel, D. S., Burrows, A., & Milsom, J. A. 2011, The
Hines, D. C. 2014, The Astrophysical Journal Letters, 791, L37
Astrophysical Journal, 727, 57
Rodigas, T. J., Males, J. R., Hinz, P. M., Mamajek, E. E., &
Knox, R. P. 2011, The Astrophysical Journal, 732, 10
Stevenson, D. J. 1991, IN: Annual review of astronomy and
astrophysics. Vol. 29 (A92-18081 05-90). Palo Alto, 29, 163
Rodigas, T. J., Weinberger, A., Mamajek, E. E., et al. 2015, ApJ,
Stolker, T., Dominik, C., Avenhaus, H., et al. 2016, arXiv,
811, 1
1603.00481v1
Rodigas, T. J., Arriagada, P., Faherty, J., et al. 2016, ApJ, 818, 1
Rodriguez, D. R., Bessell, M. S., Zuckerman, B., & Kastner, J. H.
Stone, J. M., Skemer, A. J., Kratter, K. M., et al. 2016, The
Astrophysical Journal Letters, 818, 1
2011a, The Astrophysical Journal, 727, 62
Stumpf, M. B., Brandner, W., Joergens, V., et al. 2010, The
Rodriguez, D. R., Zuckerman, B., Kastner, J. H., et al. 2013,
Astrophysical Journal, 724, 1
ApJ, 774, 101
Rodriguez, D. R., Zuckerman, B., Melis, C., & Song, I. 2011b,
The Astrophysical Journal, 732, L29
Rosotti, G. P., Juhasz, A., Booth, R. A., & Clarke, C. J. 2016,
MNRAS, 1
Su, K., Rieke, G. H., & Misselt, K. A. 2005, The Astrophysical . . .
Su, K. Y. L., Rieke, G. H., Malhotra, R., et al. 2013, ApJ, 763,
118
Sudol, J. J., & Haghighipour, N. 2012, ApJ, 755, 38
Tamura, M. 2016, Proceedings of the Japan Academy. Ser. B:
Rouan, D., Riaud, P., Boccaletti, A., Cl´enet, Y., & Labeyrie, A.
Physical and Biological Sciences, 92, 45
2000, The Publications of the Astronomical Society of the
Pacific, 112, 1479
Ryu, T., Sato, B., Kuzuhara, M., et al. 2016, arXiv, 1603.02017v1
Sallum, S., Follette, K. B., Eisner, J. A., et al. 2015a, Nature,
527, 342
Sallum, S., Eisner, J. A., Close, L. M., et al. 2015b, ApJ, 801, 85
Santos, N. C., Mayor, M., Naef, D., et al. 2002, A&A, 392, 215
Sartoretti, P., Brown, R. A., & Latham, D. W. 1998, Astronomy
and . . .
Tamura, M., SEEDS team, Usuda, T., Tamura, M., & Ishii, M.
2009, in EXOPLANETS AND DISKS: THEIR FORMATION
AND DIVERSITY: Proceedings of the International
Conference, AA(National Astronomical Observatory of Japan,
Mitaka, Tokyo 181-8588 motohide.tamuraatnao.ac.jp) (AIP),
11–16
Tanner, A., Beichman, C., Akeson, R., et al. 2007, The
Publications of the Astronomical Society of the Pacific, 119, 747
Tanner, A. M., Gelino, C. R., & Law, N. M. 2010, Publications of
Saumon, D., & Marley, M. S. 2008, The Astrophysical Journal,
the Astronomical Society of the Pacific, 122, 1195
689, 1327
Savransky, D. 2015, ApJ, 800, 100
Schlieder, J. E., L´epine, S., & Simon, M. 2010, The Astronomical
Journal, 140, 119
-. 2012a, The Astronomical Journal, 143, 80
-. 2012b, The Astronomical Journal, 144, 109
Schlieder, J. E., Skemer, A. J., Maire, A.-L., et al. 2016, ApJ,
818, 1
Schneider, A. C., Windsor, J., Cushing, M. C., Kirkpatrick, J. D.,
& Wright, E. L. 2016, The Astrophysical Journal Letters, 822, 1
Schnupp, C., Bergfors, C., Brandner, W., et al. 2010, arXiv,
astro-ph.SR
Scholz, A., Jayawardhana, R., Muzic, K., et al. 2012, ApJ, 756, 24
Scholz, R.-D. 2010, A&A, 515, A92
Schroeder, D. J., Golimowski, D. A., Brukardt, R. A., et al. 2000,
The Astronomical . . .
Serabyn, E., Mawet, D., & Burruss, R. 2010, Nature, 464, 1018
Shkolnik, E., Liu, M. C., & Reid, I. N. 2009, The Astrophysical
Journal, 699, 649
Terebey, S., van Buren, D., Matthews, K., & Padgett, D. L. 2000,
The Astronomical Journal, 119, 2341
Terebey, S., van Buren, D., Padgett, D. L., Hancock, T., &
Brundage, M. 1998, The Astrophysical Journal, 507, L71
Thalmann, C., Schmid, H. M., Boccaletti, A., et al. 2008, in SPIE
Astronomical Telescopes + Instrumentation, ed. I. S. McLean
& M. M. Casali (SPIE), 70143F
Thalmann, C., Carson, J., Janson, M., et al. 2009, The
Astrophysical Journal, 707, L123
Thalmann, C., Grady, C. A., Goto, M., et al. 2010, The
Astrophysical Journal Letters, 718, L87
Thalmann, C., Usuda, T., Kenworthy, M., et al. 2011, The
Astrophysical Journal, 732, L34
Thalmann, C., Desidera, S., Bonavita, M., et al. 2014, A&A, 572,
A91
Thomas, S., Belikov, R., & Bendek, E. 2015, ApJ, 810, 1
Tinney, C. G., Faherty, J. K., Kirkpatrick, J. D., et al. 2014, ApJ,
796, 39
31
Todorov, K., Luhman, K. L., & Mcleod, K. K. 2010, The
Wilner, D. J., Holman, M. J., & Kuchner, M. J. 2002, The
Astrophysical Journal, 714, L84
Astrophysical . . .
Todorov, K. O., Line, M. R., Pineda, J. E., et al. 2015, arXiv,
Winn, J. N., & Fabrycky, D. C. 2015, Annu. Rev. Astro.
1504.00217v3
Astrophys., 53, 409
Todorov, K. O., Luhman, K. L., Konopacky, Q. M., et al. 2014,
Winters, J. G., Henry, T. J., Lurie, J. C., et al. 2015, The
ApJ, 788, 40
Tokovinin, A. 2014, The Astronomical Journal, 147, 87
Torres, C. A. O., Quast, G. R., Melo, C. H. F., & Sterzik, M. F.
2008, in Young Nearby Loose Associations, ed. B. Reipurth
(San Francisco, CA: ASP), 757
Astronomical Journal, 149, 1
Wollert, M., Brandner, W., Reffert, S., et al. 2014, A&A, 564,
A10
Wright, J. T., Marcy, G. W., Howard, A. W., et al. 2012, ApJ,
753, 160
Torres, G. 1999, The Publications of the Astronomical Society of
Wright, J. T., Fakhouri, O., Marcy, G. W., et al. 2011,
the Pacific, 111, 169
Traub, W. A., & Oppenheimer, B. R. 2010, Exoplanets, 111
Traub, W. A., Belikov, R., Guyon, O., et al. 2014, in SPIE
Astronomical Telescopes + Instrumentation, ed. J. M.
Oschmann, M. Clampin, G. G. Fazio, & H. A. MacEwen
(SPIE), 91430N
Publications of the Astronomical Society of the Pacific, 123, 412
Wu, Y.-L., Close, L. M., Males, J. R., et al. 2015, The
Astrophysical Journal Letters, 807, 1
Wyatt, M. C. 2008, Annu. Rev. Astro. Astrophys., 46, 339
Wyatt, M. C., Holmes, E. K., Pina, R. K., et al. 1999, ApJ, 527,
918
Trilling, D. E., Bryden, G., Beichman, C. A., et al. 2008, The
Wyatt, M. C., Kennedy, G., Sibthorpe, B., et al. 2012, Monthly
Astrophysical Journal, 674, 1086
Notices RAS, 424, 1206
Troy, M., Dekany, R. G., Brack, G., et al. 2000, Proc. SPIE Vol.
Yamamoto, K., Matsuo, T., Shibai, H., et al. 2013, PASJ,
4007, 4007, 31
Vigan, A., Gry, C., Salter, G., et al. 2015, Monthly Notices RAS,
454, 129
1306.3100
Zahnle, K. J., & Marley, M. S. 2014, ApJ, 797, 41
Zapatero Osorio, M. R., Rebolo, R., Bihain, G., et al. 2010, The
Vigan, A., Langlois, M., Moutou, C., & Dohlen, K. 2008, A&A,
Astrophysical Journal, 715, 1408
489, 1345
Zhou, Y., Apai, D., Schneider, G. H., Marley, M. S., & Showman,
Vigan, A., Moutou, C., Langlois, M., et al. 2010, Monthly Notices
A. P. 2016, ApJ, 818, 1
RAS, 407, 71
Vigan, A., Patience, J., Marois, C., et al. 2012, A&A, 544, 9
Vigan, A., Bonnefoy, M., Ginski, C., et al. 2016, A&A, 587, A55
Wahhaj, Z., Koerner, D. W., Ressler, M. E., et al. 2003, The
Astrophysical Journal, 584, L27
Wahhaj, Z., Liu, M. C., Biller, B. A., et al. 2011, The
Astrophysical Journal, 729, 139
-. 2013a, ApJ, 779, 80
Wahhaj, Z., Liu, M. C., Nielsen, E. L., et al. 2013b, The
Astrophysical Journal, 773, 179
Wahhaj, Z., Liu, M. C., Biller, B. A., et al. 2014, arXiv,
1404.6525v1
Wahhaj, Z., Cieza, L. A., Mawet, D., et al. 2015, A&A, 581, A24
Wang, J., Xie, J.-W., Barclay, T., & Fischer, D. A. 2014, ApJ,
783, 4
White, R. J., & Ghez, A. M. 2001, The Astrophysical Journal,
556, 265
Williams, J. P., & Cieza, L. A. 2011, Annu. Rev. Astro.
Astrophys., 49, 67
Zhou, Y., Herczeg, G. J., Kraus, A. L., Metchev, S., & Cruz,
K. L. 2014, The Astrophysical Journal, 783, L17
Zhu, Z., Dong, R., Stone, J. M., & Rafikov, R. R. 2015, ApJ, 813,
1
Zimmerman, N., Brenner, D., Oppenheimer, B. R., et al. 2011,
Publications of the Astronomical Society of the Pacific, 123, 746
Zimmerman, N., Oppenheimer, B. R., Hinkley, S., et al. 2010,
The Astrophysical Journal, 709, 733
Zucker, S., Mazeh, T., Santos, N. C., Udry, S., & Mayor, M. 2003,
A&A, 404, 775
-. 2004, A&A, 426, 695
Zuckerman, B. 2001, Annu. Rev. Astro. Astrophys., 39, 549
Zuckerman, B., Rhee, J. H., Song, I., & Bessell, M. S. 2011, The
Astrophysical Journal, 732, 61
Zuckerman, B., & Song, I. 2004, Annu. Rev. Astro. Astrophys.,
42, 685
Zuckerman, B., Song, I., Bessell, M. S., & Webb, R. A. 2001, The
Astrophysical Journal, 562, L87
Zurlo, A., Vigan, A., Hagelberg, J., et al. 2013, A&A, 554, A21
Zurlo, A., Vigan, A., Galicher, R., et al. 2016, A&A, 587, A57
32
Directly Imaged Planets and Planet Candidates with Inferred Masses .13 MJup
TABLE 1
Name
Massa
(MJup)
Luminosity
(log (LBol/L⊙))
Age
(Myr)
Sep.
(′′)
Sep.
(AU)
NIR SpT Orbital
Pri.
Pri. Mass
References
Motion? Mult.b
(M⊙)
51 Eri b
HD 95086 b
HR 8799 b
LkCa 15 bc
HR 8799 c
HR 8799 d
HR 8799 e
β Pic b
2 ± 1
5 ± 2
5 ± 1
6 ± 4
7 ± 2
7 ± 2
7 ± 2
–5.6 ± 0.2
–4.96 ± 0.10
–5.1 ± 0.1
· · ·
–4.7 ± 0.1
–4.7 ± 0.2
–4.7 ± 0.2
12.7 ± 0.3
–3.78 ± 0.03
Close-in Planets (<100 AU)
23 ± 3
17 ± 4
40 ± 5
2 ± 1
40 ± 5
40 ± 5
40 ± 5
23 ± 3
0.45
0.6
1.7
0.08
0.95
0.62
0.38
0.4
13
56
68
20
38
24
14
9
T4.5–T6
L/T:
∼L/Tpec
· · ·
∼L/Tpec
∼L7pec
∼L7pec
L1
Planetary-Mass Companions on Wide Orbits (>100 AU)
WD 0806-661 b
Ross 458 c
ROXs 42B b
HD 106906 b
GU Psc b
CHXR 73 b
SR12 C
TYC 9486-927-1 b
7.5 ± 1.5
9 ± 3
10 ± 4
11 ± 2
11 ± 2
13 ± 6
13 ± 2
12–15
· · ·
–5.62 ± 0.03
–3.07 ± 0.07
–3.64 ± 0.08
–4.75 ± 0.15
–2.85 ± 0.14
–2.87 ± 0.20
· · ·
2000 ± 500
150–800
3 ± 2
13 ± 2
120 ± 10
2 ± 1
3 ± 2
10–45
130
102
1.2
7.1
42
1.3
8.7
217
2500
1190
140
650
2000
210
1100
6900
Y?
T8.5pec
L1
L2.5
T3.5
≥M9.5
M9.0
L3
2M1207–3932 b
2M0441+2301 Bb
5 ± 2
10 ± 2
–4.68 ± 0.05
–3.03 ± 0.09
10 ± 3
2 ± 1
0.8
41
12/0.1
1800/15
L3
L1
Planetary-Mass Companions Orbiting Brown Dwarfs
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
No
No
Yes
No
No
No
No
No
No
Yes
Candidate Planets and Companions Near the Deuterium-Burning Limit
1RXS J1609–2105 B
2M0103–5515 b
2M0122–2439 B
2M0219–3925 B
AB Pic B
CFBDSIR J1458+1013 B
DH Tau B
Fomalhaut b
FU Tau B
FW Tau b
G196-3 B
GJ 504 b
GJ 758 B
GSC 6214-210 B
HD 100546 b
HD 100546 c
HD 203030 B
HN Peg B
κ And b
LkCa 15 cc
LkCa 15 d
LP 261-75 B
ROXs12 B
SDSS2249+0044 A
SDSS2249+0044 B
VHS1256–1257 b
WISE J0146+4234 B
WISE J1217+1626 B
14 ± 2
13–35
12–27
14 ± 1
13–30
5–20
8–22
.2
∼16
∼10–100
12–25
3–30
10–40
15 ± 2
∼10 ± 5
<20
12–30
12–31
12–66
<10
<10
12–26
16 ± 4
12–60
8–52
10–21
4–16
5–20
–3.36 ± 0.09
–3.49 ± 0.11
–4.19 ± 0.10
–3.84 ± 0.05
–3.7 ± 0.2
–6.74 ± 0.19
–2.71 ± 0.12
· · ·
–2.60
· · ·
–3.8 ± 0.2
–6.13 ± 0.03
–6.1 ± 0.2
–3.1 ± 0.1
· · ·
· · ·
–4.64 ± 0.07
–4.77 ± 0.03
–3.76 ± 0.06
· · ·
· · ·
–4.43 ± 0.09
· · ·
–3.9 ± 0.3
–4.2 ± 0.3
–5.05 ± 0.22
–7.01 ± 0.22
–6.79 ± 0.18
11 ± 2
45 ± 4
120 ± 10
45 ± 4
45 ± 4
1000–5000
2 ± 1
440 ± 40
2 ± 1
2 ± 1
20–85
100–6500
1000–6000
11 ± 2
5–10
5–10
130–400
300 ± 200
40–300
2 ± 1
2 ± 1
100–200
8 ± 3
20–300
20–300
150–300
1000-10000
1000–5000
2.2
1.7
1.4
4.0
5.5
0.1
2.3
13
5.7
2.3
16
2.5
1.9
2.2
0.48
0.13
12
43
1.1
0.08
0.09
14
1.8
330
84
52
156
250
2.6
340
119
800
330
400
44
29
320
53
13
490
800
55
15
18
450
210
0.3/49
17/2600
0.3
8.1
0.09
0.08
17
102
1
8
L2
· · ·
L4
L4
L0
Y0:
M9.25
· · ·
M9.25
pec
L3
T:
T8:
M9.5
· · ·
· · ·
L7.5
T2.5
L1
· · ·
· · ·
L4.5
· · ·
L3
L5
L7
Y0
Y0
No
Yes
No
No
No
Yes
No
Yes
No
No
No
Yes
Yes
No
No
No
Yes
No
No
Yes
Yes
No
Yes
No
No
No
Yes
No
S
S
S
S
S
S
S
S
S
B
B
B
S
S
B
S
1.75
1.6
1.5
1.0
1.5
1.5
1.5
1.6
2.0d
0.6, 0.09
0.89, 0.36
1.5
0.30
0.30
1.0, 0.5
0.4
1, 2, 3
4–7
8–11
12–15
8–11
8, 10, 11
10, 11, 16
17–20
21–23
24–28
29–33
34, 35
36
37
31, 38
39, 40
S
B/S
0.024
0.2, 0.018
41–44, 11
45–47
S
B
S
S
S
S
S
S
S
B
S
S
S
S
S
S
S
S
S
S
S
S
S
S/S
S
B
S
S
0.85
0.19, 0.17
48–51
52, 53, 11
0.4
0.11
0.95
0.01–0.04
0.5
1.92
0.05
53, 54
55
56, 57
58, 59
60, 37, 15
61–64
65
0.3, 0.3
29, 31, 66
0.43
1.16
1.0
0.9
2.4
2.4
0.95
1.07
2.8
1.0
1.0
0.22
0.9
· · ·
0.03
0.07, 0.07
0.005–0.016
0.01–0.04
67–69, 53,
70–73
74–77
50, 31, 78,
80–82
83
84, 85
85, 86
87–90
14, 15
14, 15
91, 53
29, 33
92
92
93, 94
95
96
References. - (1) Macintosh et al. (2015); (2) De Rosa et al. (2015); (3) Mamajek & Bell (2014); (4) Rameau et al. (2013c); (5) Meshkat et al. (2013b
(6) De Rosa et al. (2016); (7) Rameau et al. (2016); (8) Marois et al. (2008); (9) Rajan et al. (2015); (10) Bonnefoy et al. (2016); (11) Bell et al. (2015); (12)
Kraus & Ireland (2012); (13) Ireland & Kraus (2014); (14) Sallum et al. (2015a); (15) Andrews et al. (2013); (16) Marois et al. (2010b); (17) Lagrange et
(2009a); (18) Lagrange et al. (2010); (19) Morzinski et al. (2015); (20) Bonnefoy et al. (2013); (21) Luhman et al. (2011); (22) Luhman et al. (2012); (23)
Rodriguez et al. 2011b; (24) Goldman et al. (2010); (25) Scholz (2010); (26) Burgasser et al. (2010); (27) Burningham et al. (2011); (28) Beuzit et
(2004); (29) Kraus et al. (2014a); (30) Currie et al. (2014b); (31) Bowler et al. (2014); (32) Currie et al. (2014a); (33) Bryan et al., submitted; (34)
Bailey et al. (2014); (35) Lagrange et al. (2016); (36) Naud et al. (2014); (37) Luhman et al. (2006); (38) Kuzuhara et al. (2011); (39) Deacon et al. (2016
(40) Reid et al. (2008); (41) Chauvin et al. (2004); (42) Chauvin et al. (2005a); (43) Barman et al. (2011b); (44) Allers & Liu (2013); (45) Todorov et
(2010); (46) Todorov et al. (2014); (47) Bowler & Hillenbrand (2015); (48) Lafreni`ere et al. (2008); (49) Lafreni`ere et al. (2010); (50) Ireland et al. (2011
(51) Wu et al. (2015); (52) Delorme et al. (2013); (53) Bowler et al. (2013); (54) Hinkley et al. (2015a); (55) Artigau et al. (2015); (56) Chauvin et al. (2005b
(57) Bonnefoy et al. (2010); (58) Liu et al. (2011); (59) Liu et al. (2012); (60) Itoh et al. (2005); (61) Kalas et al. (2008b); (62) Kalas et al. (2013); (63)
Mamajek (2012); (64) Janson et al. (2012); (65) Luhman et al. (2009b); (66) White & Ghez (2001); (67) Rebolo et al. (1998); (68) Zapatero Osorio et
(2010); (69) Gaidos et al. (2014); (70) Kuzuhara et al. (2013); (71) Janson et al. (2013b); (72) Fuhrmann & Chini (2015); (73) Skemer et al. (2016); (74)
Thalmann et al. (2009); (75) Currie et al. (2010); (76) Janson et al. (2011b); (77) Vigan et al. (2016); (78) Bowler et al. (2011); (79) Lachapelle et
(2015); (80) Quanz et al. (2013); (81) Currie et al. (2014d); (82) Quanz et al. (2015); (83) Currie et al. (2015); (84) Metchev & Hillenbrand (2006); (85)
Tokovinin (2014); (86) Luhman et al. (2007); (87) Carson et al. (2013); (88) Hinkley et al. (2013); (89) Bonnefoy et al. (2014a); (90) Jones et al. 2016; (91)
Reid & Walkowicz (2006); (92) Allers et al. (2010); (93) Gauza et al. (2015); (94) Stone et al. (2016); (95) Dupuy et al. (2015a); (96) Liu et al. (2012)
a Inferred masses assume hot-start evolutionary models.
b Multiplicity of the host star interior to the companion's orbit.
c LkCa 15 "b" from Kraus & Ireland (2012) is planet "c" in Sallum et al. (2015a). Here I use the original nomenclature from Kraus et al; LkCa15 c in this
table is the candidate planet "b" from Sallum et al.
Directly Imaged Stellar and Substellar Companions Inducing Shallow Radial Velocity Trends
TABLE 2
33
Name
Separation
Separation
HD 19467 B
HD 4747 B
HR 7672 B
HIP 71898 B
HD 5608 B
HD 68017 B
HD 7449 B
Gl 15 B
HD 104304 B
HD 109272 B
HD 71881 B
Kepler-444 BC
HD 164509 B
HD 126614 B
HD 197037
HAT-P-10 B
HD 41004 A
HD 77407 B
γ Cep B
HD 8375 C
HD 196885 B
HD 114174 B
HD 8049 B
HD 195109 B
Gl 86 B
Procyon B
γ Hya B
HD 53665 B
(′′)
1.65
0.60
0.5
3.0
0.6
0.59
0.54
41
1.0
1.2
0.85
1.8
0.75
0.49
3.7
0.34
0.5
1.7
0.88
0.3
0.7
0.69
1.5
2.4
2.3
4.3
1.6
0.14
(AU)
51
11.3
18
30
34
13
18
146
13
59
35
66
39
33
24
42
22
50
20
17
23
18
50
92
28
15
66
103
Nature of
Companion
BD
BD
BD
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
LMS
WD
WD
LMS
WD
WD
LMS
LMS
Comp. Mass Companion Planet
Host?
(M⊙)
SpT
dv/dt
(m s−1 yr−1) References
0.050
0.057
0.066
0.074
0.10
0.15
0.2
0.2
0.21
0.28
0.29
0.29+0.25
0.33
0.33
0.34
0.36
0.4
0.4
0.41
0.5
0.5
0.54
0.56
0.58
0.59
0.59
0.61
0.7
T5.5
· · ·
L4.5
L0
· · ·
[M5]
[M4.5]
M3.5
[M4]
· · ·
[M3]
[M]
[M]
[M]
[M]
[M]
K1
[M2]
[M4]
[M1]
M1
· · ·
· · ·
· · ·
DQ6
DQZ
· · ·
[K7]
No
No
No
No
Yes
No
Yes
Yes
No
No
No
Yes
Yes
Yes
Yes
Yes
Yes
No
Yes
No
Yes
No
No
Yes
Yes
No
No
No
–1.37 ± 0.09
–30.7 ± 0.6a
–24 ± 0.6a
8.6 ± 0.4
–5.51 ± 0.45
16.3 ± 0.9a
· · ·a
–0.27 ± 0.09
· · ·a
2.4 ± 0.2
–10.3 ± 0.2
–7.8 ± 0.5
–3.3 ± 0.5
16.2 ± 0.2
–1.9 ± 0.3
–5.1 ± 1.4
105 ± 10a
· · ·
· · ·a
52.9 ± 1.9
· · ·a
61.1 ± 0.1a
· · ·
1.9
131
· · ·a
4.1 ± 0.2
5.3 ± 0.3
1, 2
3
4, 5
6, 7, 8
9
10
11
12
13
9
10
14
15
16
17
18, 19
20, 21, 22
23
24
25
26, 27
28, 29
30
15
31–36
37, 38, 39
9
40
References. - (1) Crepp et al. (2014); (2) Crepp et al. (2015); (3) Crepp et al. (2016); (4) Liu et al. (2002); (5)
Crepp et al. (2012a); (6) Montet et al. (2014); (7) Golimowski et al. (2004b) (8) Forveille et al. (2004); (9) Ryu et al.
(2016); (10) Crepp et al. (2012b); (11) Rodigas et al. (2016); (12) Howard et al. (2014); (13) Schnupp et al. (2010); (14)
Dupuy et al. (2016); (15) Bryan et al. (2016); (16) Howard et al. (2010); (17) Ginski et al. (2016); (18) Knutson et al. (2014);
(19) Ngo et al. (2015); (20) Santos et al. (2002); (21) Zucker et al. (2003); (22) Zucker et al. (2004); (23) Mugrauer et al.
(2004); (24) Neuhauser et al. (2007); (25) Crepp et al. (2013b); (26) Chauvin et al. (2007); (27) Chauvin et al. (2011); (28)
Crepp et al. (2013a); (29) Matthews et al. (2014b); (30) Zurlo et al. (2013); (31) Queloz et al. (2000); (32) Els et al. (2001);
(33) Mugrauer & Neuhauser (2005); (34) Chauvin et al. (2006); (35) Lagrange et al. (2006); (36) Farihi et al. (2013); (37)
Irwin et al. (1992); (38) Torres (1999); (39) Bond et al. (2015); (40) Crepp et al. (2012b).
a Radial velocities show both acceleration and curvature.
34
Sample
BA
FGK
M
All Stars
The Frequency of 5–13 MJup Planets on Wide Orbits
TABLE 3
Number of Occurrence Rate Occurrence Rate Occurrence Rate Occurrence Rate
(100–1000 AU)
(10–1000 AU)
Stars
(10–100 AU)
(30–300 AU)
110
155
118
384
7.7+9.0
−6.0%
<6.8%
<4.2%
0.8+1.2
−0.5%
2.8+3.7
−2.3%
<4.1%
<3.9%
0.6+0.7
−0.5%
3.5+4.7
−2.5%
<5.8%
<5.4%
0.8+1.0
−0.6%
<6.4%
<5.1%
<7.3%
<2.1%
Note. - Assumes circular orbits, logarithmically-flat planet mass-period distributions, and hot-
start evolutionary models from Baraffe et al. (2003). All binaries within 100 AU of the host stars
have been removed. Occurrence rates are 68% credible intervals and upper limits are 95% confidence
values.
35
APPENDIX
ASTROMETRY OF HR 8799 BCDE AND β PIC B
Astrometry of HR 8799 bcde and β Pic b Until 2016
TABLE 4
UT Date
Epoch
Telescope/
Instrument
ρ
(′′)
θ
(◦)
∆R.A.
(′′)
∆Dec
(′′)
Reference
1998 Oct 30
1998 Oct 30a
2002 Jul 19
2004 Jul 14
2005 Jul 15
2007 Aug 02
2007 Aug 02a
2007 Oct 25
2007 Oct 25a
2008 Jul 11
2008 Aug 12
2008 Sep 18
2008 Sep 18a
2008 Nov 21
2008 Nov 21a
2009 Aug 15
2009 Sep 12
2009 Oct 06
2009 Oct 08
2009 Jul 11
2009 Jul 30
2009 Aug 01
2009 Nov 01
2009 Nov 01
2010 Jul 13
2010 Oct 30
2010 Oct 30a
2011 Jul 21
2011 Oct 16
2011 Nov 09
2012 Jun 14
2012 Jul 22
2012 Oct 26
2013 Oct 16
2013 Oct 21
2014 Jul 13
2014 Jul 17
2014 Dec 04–08
2014 Dec 05–09
1998 Oct 30
2004 Jul 14
2005 Jul 15
2007 Aug 02
2007 Aug 02a
2007 Oct 25
2007 Oct 25a
2008 Jul 11
2008 Aug 12
2008 Sep 18
2008 Sep 18a
2008 Nov 21
2008 Nov 21a
2009 Jan 09
2009 Jul 11
2009 Jul 30
2009 Aug 01
2009 Sep 12
2009 Sep 12a
2009 Oct 06
2009 Oct 08
2009 Nov 01
1998.828
1998.828
2002.545
2004.534
2005.534
2007.583
2007.583
2007.813
2007.813
2008.523
2008.613
2008.715
2008.715
2008.890
2008.890
2009.619
2009.695
2009.761
2009.767
2009.523
2009.575
2009.580
2009.832
2009.832
2010.528
2010.827
2010.827
2011.550
2011.789
2011.854
2012.452
2012.556
2012.819
2013.789
2013.802
2014.528
2014.539
2014.93
2014.93
1998.828
2004.534
2005.534
2007.583
2007.583
2007.813
2007.813
2008.523
2008.613
2008.715
2008.715
2008.890
2008.890
2009.022
2009.523
2009.575
2009.580
2009.695
2009.695
2009.761
2009.767
2009.832
HST /NICMOS
HST /NICMOS
Subaru/CIAO
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Gemini-N/NIRI
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
MMT/Clio
MMT/Clio
Subaru/IRCS
MMT/Clio
VLT/NaCo
VLT/NaCo
Palomar/WCS
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/PISCES
LBT/PISCES
Palomar/P1640
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/LMIRCam
VLT/SPHERE
Keck/NIRC2
VLT/SPHERE
VLT/SPHERE
HST /NICMOS
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Gemini-N/NIRI
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
MMT/Clio
MMT/Clio
MMT/Clio
Palomar/WCS
Keck/NIRC2
Keck/NIRC2
MMT/Clio
MMT/Clio
VLT/NaCo
VLT/NaCo
Keck/NIRC2
1.721 ± 0.012
1.738 ± 0.018
[1.74 ± 0.03]
[1.716 ± 0.007]
[1.724 ± 0.007]
[1.726 ± 0.004]
[1.721 ± 0.004]
[1.713 ± 0.007]
[1.717 ± 0.010]
[1.723 ± 0.006]
[1.724 ± 0.003]
[1.724 ± 0.004]
[1.723 ± 0.006]
[1.728 ± 0.017]
[1.73 ± 0.03]
[1.725 ± 0.014]
[1.72 ± 0.04]
[1.74 ± 0.03]
[1.720 ± 0.010]
1.73 ± 0.10
[1.721 ± 0.006]
[1.725 ± 0.010]
[1.74 ± 0.03]
[1.719 ± 0.014]
[1.720 ± 0.007]
[1.716 ± 0.021]
[1.717 ± 0.007]
[1.719 ± 0.007]
[1.741 ± 0.016]
[1.708 ± 0.016]
[1.715 ± 0.007]
[1.716 ± 0.007]
[1.718 ± 0.006]
[1.706 ± 0.031]
1.717 ± 0.013
[1.722 ± 0.008]
[1.720 ± 0.018]
[1.724 ± 0.007]
1.729 ± 0.007
0.966 ± 0.022
[0.960 ± 0.007]
[0.951 ± 0.007]
[0.952 ± 0.007]
[0.957 ± 0.006]
[0.958 ± 0.007]
[0.957 ± 0.010]
[0.961 ± 0.006]
[0.964 ± 0.003]
[0.964 ± 0.004]
[0.959 ± 0.004]
[0.921 ± 0.03]
[0.958 ± 0.03]
[0.90 ± 0.04]
1.00 ± 0.06
[0.957 ± 0.006]
[0.962 ± 0.013]
[0.96 ± 0.03]
[0.942 ± 0.04]
[0.940 ± 0.06]
[0.952 ± 0.010]
[0.957 ± 0.018]
HR 8799 b
55.1 ± 0.4
54.7 ± 0.4
[58.2 ± 0.6]
[59.0 ± 0.17]
[60.2 ± 0.17]
[61.8 ± 0.1]
[60.90 ± 0.10]
[62.0 ± 0.17]
[60.9 ± 0.2]
[62.4 ± 0.13]
[62.3 ± 0.07]
[62.4 ± 0.10]
[61.65 ± 0.13]
[63.2 ± 0.4]
[62.5 ± 0.7]
[62.9 ± 0.3]
[63.2 ± 1.0]
[62.0 ± 0.7]
[62.9 ± 0.2]
62.9 ± 1.5
[62.42 ± 0.13]
[62.6 ± 0.2]
[62.5 ± 0.6]
[62.5 ± 0.3]
[62.93 ± 0.17]
[63.5 ± 0.5]
[64.2 ± 0.17]
[63.69 ± 0.17]
[65.1 ± 0.4]
[64.9 ± 0.4]
[65.7 ± 0.17]
[64.20 ± 0.17]
[64.37 ± 0.13]
[64.9 ± 0.7]
65.46 ± 0.44
[65.8 ± 1.7]
[65.1 ± 0.4]
[65.9 ± 0.17]
65.93 ± 0.15
HR 8799 c
300 ± 0.8
[309.6 ± 0.3]
[311.5 ± 0.3]
[315.1 ± 0.3]
[314.5 ± 0.2]
[315.3 ± 0.3]
[314.9 ± 0.4]
[316.8 ± 0.2]
[317.06 ± 0.12]
[317.06 ± 0.18]
[316.27 ± 0.18]
[316.8 ± 1.1]
[316.9 ± 1.2]
[317.4 ± 1.9]
315.9 ± 1.5
[318.1 ± 0.2]
[318.7 ± 0.5]
[319.2 ± 1.2]
[317.7 ± 1.8]
[317 ± 2]
[318.8 ± 0.4]
[318.8 ± 0.8]
1.411 ± 0.009
[1.418 ± 0.016]
1.481 ± 0.023
1.471 ± 0.005
1.496 ± 0.005
1.522 ± 0.003
1.504 ± 0.003
1.512 ± 0.005
1.500 ± 0.007
1.527 ± 0.004
1.527 ± 0.002
1.528 ± 0.003
1.516 ± 0.004
1.54 ± 0.01
1.532 ± 0.02
1.536 ± 0.01
1.538 ± 0.03
1.535 ± 0.02
1.532 ± 0.007
1.54 ± 0.04
1.526 ± 0.004
1.531 ± 0.007
1.54 ± 0.019
1.524 ± 0.010
1.532 ± 0.005
1.535 ± 0.015
1.546 ± 0.005
1.541 ± 0.005
1.579 ± 0.011
1.546 ± 0.011
1.563 ± 0.005
1.545 ± 0.005
1.549 ± 0.004
1.545 ± 0.022
1.562 ± 0.008
1.570 ± 0.006
1.560 ± 0.013
1.574 ± 0.005
[1.579 ± 0.007]
[–0.84 ± 0.02]
–0.739 ± 0.005
–0.713 ± 0.005
–0.672 ± 0.005
–0.683 ± 0.004
–0.674 ± 0.005
–0.678 ± 0.007
–0.658 ± 0.004
–0.657 ± 0.002
–0.657 ± 0.003
–0.663 ± 0.003
–0.631 ± 0.015
–0.654 ± 0.02
–0.612 ± 0.03
–0.70 ± 0.05
–0.639 ± 0.004
–0.635 ± 0.009
–0.625 ± 0.02
–0.634 ± 0.03
–0.636 ± 0.04
–0.627 ± 0.007
–0.63 ± 0.013
0.986 ± 0.009
[1.00 ± 0.014]
0.919 ± 0.017
0.884 ± 0.005
0.856 ± 0.005
0.815 ± 0.003
0.837 ± 0.003
0.805 ± 0.005
0.836 ± 0.007
0.799 ± 0.004
0.801 ± 0.002
0.798 ± 0.003
0.818 ± 0.004
0.780 ± 0.014
0.796 ± 0.02
0.785 ± 0.01
0.777 ± 0.03
0.816 ± 0.02
0.783 ± 0.007
0.79 ± 0.06
0.797 ± 0.004
0.794 ± 0.007
0.80 ± 0.019
0.795 ± 0.010
0.783 ± 0.005
0.766 ± 0.015
0.748 ± 0.005
0.762 ± 0.005
0.734 ± 0.011
0.725 ± 0.011
0.706 ± 0.005
0.747 ± 0.005
0.743 ± 0.004
0.724 ± 0.022
0.713 ± 0.013
0.707 ± 0.006
0.725 ± 0.013
0.703 ± 0.005
[0.705 ± 0.007]
[0.483 ± 0.016]
0.612 ± 0.005
0.630 ± 0.005
0.674 ± 0.005
0.671 ± 0.004
0.681 ± 0.005
0.676 ± 0.007
0.701 ± 0.004
0.706 ± 0.002
0.706 ± 0.003
0.693 ± 0.003
0.671 ± 0.02
0.700 ± 0.02
0.665 ± 0.03
0.72 ± 0.05
0.712 ± 0.004
0.722 ± 0.009
0.725 ± 0.02
0.697 ± 0.03
0.692 ± 0.04
0.716 ± 0.007
0.72 ± 0.013
Lafreni`ere et al. (2009
Soummer et al. (2011
Fukagawa et al. (2009
Marois et al. (2008
Currie et al. (2012a
Metchev et al. (2009
Konopacky et al. (2016
Marois et al. (2008
Konopacky et al. (2016
Marois et al. (2008
Marois et al. (2008
Marois et al. (2008
Konopacky et al. (2016
Hinz et al. (2010)
Currie et al. (2011b
Currie et al. (2011b
Currie et al. (2011b
Bergfors et al. (2011
Currie et al. (2011b
Serabyn et al. (2010
Konopacky et al. (2016
Konopacky et al. (2016
Galicher et al. (2011
Konopacky et al. (2016
Konopacky et al. (2016
Konopacky et al. (2016
Currie et al. (2012a
Konopacky et al. (2016
Esposito et al. (2012
Esposito et al. (2012
Pueyo et al. (2015)
Konopacky et al. (2016
Konopacky et al. (2016
Konopacky et al. (2016
Maire et al. (2015a
Zurlo et al. (2016)
Konopacky et al. (2016
Zurlo et al. (2016)
Apai et al. (2016)
Soummer et al. (2011
Marois et al. (2008
Currie et al. (2012a
Metchev et al. (2009
Konopacky et al. (2016
Marois et al. (2008
Konopacky et al. (2016
Marois et al. (2008
Marois et al. (2008
Marois et al. (2008
Konopacky et al. (2016
Hinz et al. (2010)
Currie et al. (2011b
Hinz et al. (2010)
Serabyn et al. (2010
Konopacky et al. (2016
Konopacky et al. (2016
Hinz et al. (2010)
Currie et al. (2011b
Bergfors et al. (2011
Currie et al. (2011b
Galicher et al. (2011
36
UT Date
Epoch
2009 Nov 01
2010 Jul 13
2010 Oct 30
2010 Oct 30
2011 Jul 21
2011 Oct 16
2011 Nov 09
2012 Jun 14
2012 Jul 22
2012 Oct 26
2013 Oct 16
2013 Oct 21
2014 Jul 13
2014 Jul 17
2014 Dec 04–08
2014 Dec 05–09
1998 Oct 30
2005 Jul 15
2007 Aug 02
2007 Aug 02a,b
2007 Oct 25a
2008 Jul 11
2008 Aug 12
2008 Sep 18
2008 Sep 18a
2008 Nov 21
2008 Nov 21a
2009 Jul 11
2009 Jul 30
2009 Aug 01
2009 Sep 12
2009 Oct 06
2009 Oct 08
2009 Nov 01
2009 Nov 01
2010 Jul 13
2010 Oct 30
2010 Oct 30a
2011 Jul 21
2011 Oct 16
2011 Nov 09
2012 Jun 14
2012 Jul 22
2013 Oct 16
2013 Oct 21
2012 Oct 26
2014 Jul 13
2014 Jul 17
2014 Aug 12
2014 Dec 04–08
2014 Dec 05–09
2009 Jul 30
2009 Jul 31
2009 Aug 01
2009 Aug 01a
2009 Oct 08
2009 Nov 01
2009 Nov 01a
2010 Jul 13
2010 Jul 13a
2010 Jul 21
2010 Oct 30
2010 Oct 30a
2011 Jul 21
2011 Oct 16
2011 Nov 09
2012 Jun 14
2012 Jul 22
2012 Oct 26
2009.832
2010.528
2010.827
2010.827a
2011.550
2011.789
2011.854
2012.452
2012.556
2012.819
2013.789
2013.802
2014.528
2014.539
2014.93
2014.93
1998.828
2005.534
2007.583
2007.583
2007.813
2008.523
2008.613
2008.715
2008.715
2008.890
2008.890
2009.523
2009.575
2009.580
2009.695
2009.761
2009.767
2009.832
2009.832
2010.528
2010.827
2010.827
2011.550
2011.789
2011.854
2012.452
2012.556
2013.789
2013.802
2012.819
2014.528
2014.539
2014.611
2014.93
2014.93
2009.575
2009.578
2009.580
2009.580
2009.767
2009.832
2009.832
2010.528
2010.528
2010.550
2010.827
2010.827
2011.550
2011.789
2011.854
2012.452
2012.556
2012.819
Telescope/
Instrument
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/PISCES
LBT/PISCES
Palomar/P1640
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/LMIRCam
VLT/SPHERE
Keck/NIRC2
VLT/SPHERE
VLT/SPHERE
HST /NICMOS
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
MMT/Clio
MMT/Clio
Palomar/WCS
Keck/NIRC2
Keck/NIRC2
MMT/Clio
VLT/NaCo
VLT/NaCo
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/PISCES
LBT/PISCES
Palomar/P1640
Keck/NIRC2
Keck/NIRC2
LBT/LMIRCam
Keck/NIRC2
VLT/SPHERE
Keck/NIRC2
VLT/SPHERE
VLT/SPHERE
VLT/SPHERE
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
VLT/NaCo
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
Keck/NIRC2
LBT/PISCES
LBT/PISCES
Palomar/P1640
Keck/NIRC2
Keck/NIRC2
TABLE 4 - Continued
ρ
(′′)
θ
(◦)
∆R.A.
(′′)
∆Dec
(′′)
Reference
[0.961 ± 0.013]
[0.956 ± 0.006]
[0.960 ± 0.017]
[0.949 ± 0.007]
[0.955 ± 0.006]
[0.938 ± 0.014]
[0.960 ± 0.014]
[0.947 ± 0.006]
[0.956 ± 0.007]
[0.958 ± 0.004]
[0.953 ± 0.031]
0.951 ± 0.014
[0.948 ± 0.006]
[0.964 ± 0.018]
[0.951 ± 0.006]
0.947 ± 0.007
0.549 ± 0.028
[0.585 ± 0.014]
[0.613 ± 0.011]
[0.615 ± 0.007]
[0.614 ± 0.014]
[0.618 ± 0.006]
[0.621 ± 0.003]
[0.621 ± 0.004]
[0.622 ± 0.006]
[0.68 ± 0.02]
[0.646 ± 0.03]
0.64 ± 0.04
[0.624 ± 0.004]
[0.622 ± 0.010]
[0.65 ± 0.04]
[0.66 ± 0.10]
[0.634 ± 0.010]
[0.63 ± 0.02]
[0.626 ± 0.010]
[0.634 ± 0.006]
[0.634 ± 0.018]
[0.634 ± 0.007]
[0.638 ± 0.007]
[0.637 ± 0.014]
[0.635 ± 0.014]
[0.620 ± 0.008]
[0.650 ± 0.007]
[0.647 ± 0.023]
0.657 ± 0.013
[0.648 ± 0.006]
[0.658 ± 0.008]
[0.667 ± 0.016]
[0.658 ± 0.006]
[0.660 ± 0.006]
0.654 ± 0.007
[0.372 ± 0.010]
[0.369 ± 0.03]
[0.368 ± 0.018]
[0.373 ± 0.014]
[0.375 ± 0.010]
[0.362 ± 0.014]
[0.362 ± 0.013]
[0.368 ± 0.011]
[0.363 ± 0.008]
[0.368 ± 0.016]
[0.371 ± 0.014]
[0.370 ± 0.027]
[0.375 ± 0.011]
[0.347 ± 0.016]
[0.403 ± 0.016]
[0.377 ± 0.008]
[0.382 ± 0.011]
[0.378 ± 0.013]
[318.5 ± 0.5]
[319.6 ± 0.2]
[320.8 ± 0.7]
[320.9 ± 0.3]
[321.5 ± 0.2]
[323.3 ± 0.6]
[323.0 ± 0.6]
[323.9 ± 0.2]
[322.8 ± 0.3]
[323.32 ± 0.18]
[325.3 ± 1.3]
325.5 ± 0.9
[326.6 ± 0.2]
[325.9 ± 0.8]
[327.0 ± 0.2]
327.8 ± 0.4
HR 8799 d
166 ± 3
[188.6 ± 1.0]
[196.1 ± 0.7]
[196.9 ± 0.5]
[196.5 ± 0.9]
[199.7 ± 0.4]
[200.36 ± 0.18]
[200.4 ± 0.3]
[199.0 ± 0.4]
[198.9 ± 1.7]
[199.6 ± 1.8]
197.5 ± 1.5
[202.3 ± 0.3]
[203.7 ± 0.6]
[205 ± 3]
[204 ± 6]
[202.4 ± 0.6]
[202.5 ± 1.3]
[203.7 ± 0.6]
[204.7 ± 0.4]
[207.8 ± 1.2]
[206.5 ± 0.6]
[208.3 ± 0.4]
[208.0 ± 0.9]
[210.2 ± 0.9]
[211.4 ± 0.6]
[211.4 ± 0.4]
[216.2 ± 1.4]
215.0 ± 1.2
[212.3 ± 0.4]
[216.3 ± 0.5]
[216.8 ± 0.9]
[216.5 ± 0.3]
[217.5 ± 0.3]
217.6 ± 0.8
HR 8799 e
[235.4 ± 1.1]
[234 ± 3]
[235 ± 2]
[238.5 ± 1.5]
[234.7 ± 1.1]
[237.2 ± 1.6]
[238.9 ± 1.4]
[242.0 ± 1.3]
[242.8 ± 1.0]
[241.6 ± 1.7]
[244.1 ± 1.5]
[247.2 ± 2.5]
[249.7 ± 1.2]
[249.9 ± 1.8]
[251.6 ± 1.6]
[256.2 ± 0.9]
[257.3 ± 1.2]
[258.4 ± 1.4]
–0.636 ± 0.009
–0.619 ± 0.004
–0.607 ± 0.012
–0.598 ± 0.005
–0.595 ± 0.004
–0.561 ± 0.010
–0.578 ± 0.010
–0.558 ± 0.004
–0.578 ± 0.005
–0.572 ± 0.003
–0.542 ± 0.022
–0.538 ± 0.006
–0.522 ± 0.004
–0.540 ± 0.013
–0.518 ± 0.004
[–0.504 ± 0.007]
[0.13 ± 0.03]
–0.087 ± 0.010
–0.170 ± 0.008
–0.179 ± 0.005
–0.175 ± 0.010
–0.208 ± 0.004
–0.216 ± 0.002
–0.216 ± 0.003
–0.202 ± 0.004
–0.22 ± 0.021
–0.217 ± 0.02
–0.19 ± 0.07
–0.237 ± 0.003
–0.250 ± 0.007
–0.28 ± 0.03
–0.270 ± 0.07
–0.241 ± 0.007
–0.24 ± 0.014
–0.251 ± 0.007
–0.265 ± 0.004
–0.296 ± 0.013
–0.283 ± 0.005
–0.303 ± 0.005
–0.299 ± 0.010
–0.320 ± 0.010
–0.323 ± 0.006
–0.339 ± 0.005
–0.382 ± 0.016
–0.377 ± 0.007
–0.346 ± 0.004
–0.390 ± 0.005
–0.400 ± 0.011
–0.391 ± 0.004
–0.402 ± 0.004
[–0.400 ± 0.009]
–0.306 ± 0.007
–0.299 ± 0.019
–0.303 ± 0.013
–0.318 ± 0.010
–0.306 ± 0.007
–0.304 ± 0.010
–0.310 ± 0.009
–0.325 ± 0.008
–0.323 ± 0.006
–0.324 ± 0.011
–0.334 ± 0.010
–0.341 ± 0.016
–0.352 ± 0.008
–0.326 ± 0.011
–0.382 ± 0.011
–0.366 ± 0.006
–0.373 ± 0.008
–0.370 ± 0.009
0.720 ± 0.009
0.728 ± 0.004
0.744 ± 0.012
0.737 ± 0.005
0.747 ± 0.004
0.752 ± 0.010
0.767 ± 0.010
0.765 ± 0.004
0.761 ± 0.005
0.768 ± 0.003
0.784 ± 0.022
0.784 ± 0.013
0.791 ± 0.004
0.799 ± 0.013
0.797 ± 0.004
[0.802 ± 0.007]
[–0.53 ± 0.03]
–0.578 ± 0.010
–0.589 ± 0.008
–0.588 ± 0.005
–0.589 ± 0.010
–0.582 ± 0.004
–0.582 ± 0.002
–0.582 ± 0.003
–0.588 ± 0.004
–0.644 ± 0.013
–0.608 ± 0.02
–0.61 ± 0.03
–0.577 ± 0.003
–0.570 ± 0.007
–0.59 ± 0.03
–0.600 ± 0.07
–0.586 ± 0.007
–0.58 ± 0.014
–0.573 ± 0.007
–0.576 ± 0.004
–0.561 ± 0.013
–0.567 ± 0.005
–0.562 ± 0.005
–0.563 ± 0.010
–0.549 ± 0.010
–0.529 ± 0.006
–0.555 ± 0.005
–0.522 ± 0.016
–0.538 ± 0.011
–0.548 ± 0.004
–0.530 ± 0.006
–0.534 ± 0.011
–0.529 ± 0.004
–0.523 ± 0.004
[–0.518 ± 0.008]
–0.211 ± 0.007
–0.217 ± 0.019
–0.209 ± 0.013
–0.195 ± 0.010
–0.217 ± 0.007
–0.196 ± 0.010
–0.187 ± 0.009
–0.173 ± 0.008
–0.166 ± 0.006
–0.175 ± 0.011
–0.162 ± 0.010
–0.143 ± 0.016
–0.130 ± 0.008
–0.119 ± 0.011
–0.127 ± 0.011
–0.090 ± 0.006
–0.084 ± 0.008
–0.076 ± 0.009
Konopacky et al. (2016
Konopacky et al. (2016
Konopacky et al. (2016
Currie et al. (2012a
Konopacky et al. (2016
Esposito et al. (2012
Esposito et al. (2012
Pueyo et al. (2015)
Konopacky et al. (2016
Konopacky et al. (2016
Konopacky et al. (2016
Maire et al. (2015a
Zurlo et al. (2016)
Konopacky et al. (2016
Zurlo et al. (2016)
Apai et al. (2016)
Soummer et al. (2011
Currie et al. (2012a
Metchev et al. (2009
Konopacky et al. (2016
Konopacky et al. (2016
Marois et al. (2008
Marois et al. (2008
Marois et al. (2008
Konopacky et al. (2016
Hinz et al. (2010)
Currie et al. (2011b
Serabyn et al. (2010
Konopacky et al. (2016
Konopacky et al. (2016
Hinz et al. (2010)
Bergfors et al. (2011
Currie et al. (2011b
Galicher et al. (2011
Konopacky et al. (2016
Konopacky et al. (2016
Konopacky et al. (2016
Currie et al. (2012a
Konopacky et al. (2016
Esposito et al. (2012
Esposito et al. (2012
Pueyo et al. (2015)
Konopacky et al. (2016
Konopacky et al. (2016
Maire et al. (2015a
Konopacky et al. (2016
Zurlo et al. (2016)
Konopacky et al. (2016
Zurlo et al. (2016)
Zurlo et al. (2016)
Apai et al. (2016)
Konopacky et al. (2016
Marois et al. (2010b
Marois et al. (2010b
Konopacky et al. (2016
Currie et al. (2011b
Marois et al. (2010b
Konopacky et al. (2016
Marois et al. (2010b
Konopacky et al. (2016
Marois et al. (2010b
Marois et al. (2010b
Konopacky et al. (2016
Konopacky et al. (2016
Esposito et al. (2012
Esposito et al. (2012
Pueyo et al. (2015)
Konopacky et al. (2016
Konopacky et al. (2016
TABLE 4 - Continued
37
UT Date
Epoch
2013 Oct 16
2013 Oct 21
2014 Jul 13
2014 Jul 17
2014 Aug 12
2014 Dec 04–08
2014 Dec 05–09
2013.789
2013.802
2014.528
2014.539
2014.611
2014.93
2014.93
Telescope/
Instrument
Keck/NIRC2
LBT/LMIRCam
VLT/SPHERE
Keck/NIRC2
VLT/SPHERE
VLT/SPHERE
VLT/SPHERE
ρ
(′′)
θ
(◦)
∆R.A.
(′′)
∆Dec
(′′)
Reference
[0.373 ± 0.018]
0.395 ± 0.012
[0.386 ± 0.012]
[0.387 ± 0.016]
[0.384 ± 0.003]
[0.384 ± 0.014]
0.381 ± 0.007
[267.4 ± 2.0]
264.8 ± 1.7
[268.8 ± 1.3]
[270.4 ± 1.7]
[269.3 ± 0.3]
[272.1 ± 1.5]
272.5 ± 0.7
–0.373 ± 0.013
–0.394 ± 0.011
–0.386 ± 0.009
–0.387 ± 0.011
–0.384 ± 0.002
–0.384 ± 0.010
[–0.381 ± 0.007]
–0.017 ± 0.013
–0.036 ± 0.017
–0.008 ± 0.009
0.003 ± 0.011
–0.005 ± 0.002
0.014 ± 0.010
[0.016 ± 0.005]
Konopacky et al. (2016
Maire et al. (2015a
Zurlo et al. (2016)
Konopacky et al. (2016
Zurlo et al. (2016)
Zurlo et al. (2016)
Apai et al. (2016)
2003 Nov 10
2003 Nov 10
2003 Nov 10a
2003 Nov 13
2008 Nov 11
2009 Oct 25
2009 Oct 25a
2009 Nov–Decc
2009 Dec 03
2009 Dec 03
2009 Dec 26
2009 Dec 26a
2009 Dec 29
2009 Dec 29a
2009 Dec 29a
2010 Mar 20d
2010 Apr 03
2010 Apr 10d
2010 Apr 10a
2010 Sep 28
2010 Sep 28a
2010 Nov 16
2010 Nov 17
2010 Dec 25
2010 Dec 25a
2011 Jan 02
2011 Mar 26
2011 Oct 12
2011 Oct 20
2011 Oct 20
2011 Dec 11
2012 Mar 29
2012 Mar 29
2012 Nov 01
2012 Dec 04
2012 Dec 01–07
2012 Dec 16
2013 Jan 31
2013 Jan 31a
2013 Nov 16
2013 Nov 16
2013 Nov 18
2013 Nov 18a
2013 Dec 10
2013 Dec 10a
2013 Dec 10
2013 Dec 11
2013 Dec 12
2014 Mar 23
2014 Nov 08
2015 Jan 24
2009.9
VLT/NaCo+APP
Gemini-S/NICI
Gemini-S/NICI
Gemini-S/NICI
Gemini-S/NICI
2003.857
2003.857
2003.857
2003.865
2008.862
2009.813
2009.813
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
VLT/NaCo
0.411 ± 0.008
0.411 ± 0.008
0.413 ± 0.022
0.401 ± 0.008
0.210 ± 0.027
0.298 ± 0.016
0.299 ± 0.014
0.299 ± 0.016
0.323 ± 0.010
2009.920
0.339 ± 0.010
2009.920
0.302 ± 0.016
2009.983
0.299 ± 0.013
2009.983
0.314 ± 0.016
2009.991
0.326 ± 0.013
2009.991
0.306 ± 0.009
2009.991
0.345 ± 0.007
2010.214
0.354 ± 0.012
2010.252
0.355 ± 0.002
2010.271
0.346 ± 0.007
2010.271
0.383 ± 0.011
2010.739
0.385 ± 0.011
2010.739
0.387 ± 0.008
2010.873
0.390 ± 0.013
2010.876
0.404 ± 0.010
2010.980
0.407 ± 0.005
2010.980
0.408 ± 0.009
2011.003
0.426 ± 0.013
2011.230
0.439 ± 0.005
2011.778
0.455 ± 0.003
2011.800
0.452 ± 0.005
2011.800
0.441 ± 0.003
2011.942
0.447 ± 0.003
2012.341
0.448 ± 0.005
2012.341
0.456 ± 0.011
2012.835
0.470 ± 0.010
2012.925
0.461 ± 0.014
2012.92
2012.958
0.449 ± 0.006
2013.082 VLT/NaCo+AGPM 0.452 ± 0.010
2013.082 VLT/NaCo+AGPM 0.448 ± 0.004
0.430 ± 0.003
2013.873
2013.873
0.426 ± 0.003
0.434 ± 0.006
2013.879
0.428 ± 0.003
2013.879
0.430 ± 0.010
2013.939
0.419 ± 0.006
2013.939
0.419 ± 0.004
2013.939
2013.939
0.419 ± 0.005
0.427 ± 0.007
2013.945
0.413 ± 0.003
2014.222
0.363 ± 0.004
2014.852
2015.063
0.348 ± 0.005
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Gemini-S/GPI
Magellan/MagAO
Magellan/MagAO
VLT/NaCo
VLT/NaCo
VLT/NaCo
Gemini-S/NICI
Gemini-S/NICI
VLT/NaCo
Gemini-S/NICI
Gemini-S/NICI
VLT/NaCo
VLT/NaCo
β Pic b
31.8 ± 1.3
31.5 ± 1.3
34.4 ± 3.5
32.1 ± 1.4
211.5 ± 1.9
210.6 ± 3.6
210.7 ± 2.9
209.4 ± 3.5
209.3 ± 1.8
209.2 ± 1.7
212.8 ± 3.7
211.1 ± 1.7
211.5 ± 3.5
210.6 ± 1.2
212.1 ± 1.7
209.4 ± 0.3
209.1 ± 2.1
208.4 ± 1.1
209.9 ± 1.2
210.3 ± 1.7
210.1 ± 1.5
212.4 ± 1.4
212.3 ± 2.1
212.1 ± 0.7
212.9 ± 1.4
211.1 ± 1.5
210.1 ± 1.8
212.9 ± 0.5
211.9 ± 0.4
211.6 ± 0.6
212.9 ± 0.3
210.8 ± 0.4
211.8 ± 0.6
211.9 ± 2.3
212.0 ± 1.2
211.9 ± 1.2
211.6 ± 0.7
211.2 ± 1.3
212.3 ± 0.3
212.3 ± 0.4
212.8 ± 0.4
211.8 ± 0.5
212.2 ± 0.4
211.6 ± 1.3
212.2 ± 0.8
212.6 ± 0.5
212.3 ± 0.7
211.8 ± 0.7
212.1 ± 0.4
212.2 ± 0.7
212.2 ± 0.7
[0.217 ± 0.009]
[0.215 ± 0.009]
[0.233 ± 0.022]
[0.213 ± 0.009]
[–0.110 ± 0.015]
[–0.152 ± 0.018]
–0.153 ± 0.014
[–0.147 ± 0.018]
[–0.158 ± 0.010]
[–0.165 ± 0.010]
[–0.164 ± 0.019]
[–0.154 ± 0.010]
[–0.164 ± 0.018]
[–0.166 ± 0.009]
–0.163 ± 0.009]
[–0.169 ± 0.004]
[–0.172 ± 0.013]
[–0.169 ± 0.006]
–0.173 ± 0.007
–0.193 ± 0.011
[–0.193 ± 0.010]
–0.207 ± 0.008
–0.209 ± 0.013
[–0.215 ± 0.007]
[–0.221 ± 0.009]
–0.211 ± 0.009
–0.214 ± 0.012
[–0.238 ± 0.003]
[–0.240 ± 0.003]
[–0.237 ± 0.005]
[–0.240 ± 0.003]
[–0.229 ± 0.003]
[–0.236 ± 0.005]
–0.241 ± 0.013
[–0.249 ± 0.010]
[–0.244 ± 0.011]
[–0.235 ± 0.006]
[–0.234 ± 0.010]
[–0.239 ± 0.003]
[–0.230 ± 0.003]
[–0.231 ± 0.003]
[–0.229 ± 0.005]
[–0.228 ± 0.003]
[–0.225 ± 0.010]
[–0.223 ± 0.006]
[–0.226 ± 0.004]
[–0.224 ± 0.005]
[–0.225 ± 0.006]
[–0.219 ± 0.003]
[–0.193 ± 0.004]
[–0.185 ± 0.004]
Milli et al. (2014)
Quanz et al. (2010
Lagrange et al. (2009a
Lagrange et al. (2009a
Chauvin et al. (2012
Lagrange et al. (2009a
Lagrange et al. (2010
Currie et al. (2011a
Chauvin et al. (2012
Bonnefoy et al. (2011
Bonnefoy et al. (2011
Chauvin et al. (2012
Chauvin et al. (2012
Milli et al. (2014)
Chauvin et al. (2012
Chauvin et al. (2012
Boccaletti et al. (2013
Currie et al. (2011a
Lagrange et al. (2010
Chauvin et al. (2012
Lagrange et al. (2010
Nielsen et al. (2014
Nielsen et al. (2014
Lagrange et al. (2010
[0.349 ± 0.008]
[0.350 ± 0.008]
0.341 ± 0.022
[0.340 ± 0.009]
[–0.179 ± 0.023]
[–0.257 ± 0.017]
–0.257 ± 0.014
[–0.260 ± 0.017]
[–0.282 ± 0.010]
[–0.296 ± 0.010]
[–0.254 ± 0.017]
[–0.256 ± 0.012]
[–0.268 ± 0.017]
[–0.280 ± 0.012]
[–0.260 ± 0.008
[–0.301 ± 0.006]
[–0.309 ± 0.012]
[–0.312 ± 0.004]
–0.300 ± 0.007
–0.331 ± 0.011
[–0.333 ± 0.011]
–0.326 ± 0.010
–0.330 ± 0.014
[–0.342 ± 0.009]
[–0.342 ± 0.007]
–0.350 ± 0.010
–0.367 ± 0.014
[–0.369 ± 0.003]
[–0.386 ± 0.003]
[–0.385 ± 0.005]
[–0.370 ± 0.003]
[–0.384 ± 0.003]
[–0.381 ± 0.005]
–0.387 ± 0.014
[–0.399 ± 0.010]
[–0.391 ± 0.013]
Milli et al. (2014)
[–0.382 ± 0.006]
Absil et al. (2013)
[–0.387 ± 0.010]
[–0.379 ± 0.004]
Milli et al. (2014)
[–0.363 ± 0.003] Millar-Blanchaer et al. (
[–0.358 ± 0.003] Millar-Blanchaer et al. (
[–0.369 ± 0.005]
[–0.362 ± 0.003] Millar-Blanchaer et al. (
[–0.366 ± 0.010]
[–0.355 ± 0.006] Millar-Blanchaer et al. (
[–0.353 ± 0.004] Millar-Blanchaer et al. (
[–0.354 ± 0.005] Millar-Blanchaer et al. (
[–0.363 ± 0.007] Millar-Blanchaer et al. (
[–0.350 ± 0.003] Millar-Blanchaer et al. (
[–0.307 ± 0.004] Millar-Blanchaer et al. (
[–0.294 ± 0.005] Millar-Blanchaer et al. (
Nielsen et al. (2014
Chauvin et al. (2012
Chauvin et al. (2012
Milli et al. (2014)
Nielsen et al. (2014
Nielsen et al. (2014
Milli et al. (2014)
Nielsen et al. (2014
Nielsen et al. (2014
Bonnefoy et al. (2013
Males et al. (2014)
Macintosh et al. (2014
Morzinski et al. (2015
Bonnefoy et al. (2014b
Note. - Astrometry listed in brackets either in ρ and θ or ∆R.A. and ∆Dec is inferred from published measurements.
a
b
Re-reduction of published data.
Konopacky et al. (2016) note that this epoch of astrometry for HR 8799 d may be biased because the primary is positioned near the edge of
the coronagraph.
c
d
methods.
Sum of observations taken on 2009 Nov 24, Nov 25, and Dec 17.
Bonnefoy et al. (2011) report astrometry using two reduction procedures for each data set. Values listed here are the weighted average of both
|
1909.04884 | 2 | 1909 | 2019-09-15T16:04:34 | The potassium absorption on HD189733b and HD209458b | [
"astro-ph.EP"
] | In this work, we investigate the potassium excess absorption around 7699A of the exoplanets HD189733b and HD209458b. For this purpose, we used high spectral resolution transit observations acquired with the 2 x 8.4m Large Binocular Telescope (LBT) and the Potsdam Echelle Polarimetric and Spectroscopic Instrument (PEPSI). For a bandwidth of 0.8A, we present a detection > 7-sigma with an absorption level of 0.18% for HD189733b. Applying the same analysis to HD209458b, we can set 3-sigma upper limit of 0.09%, even though we do not detect a K- excess absorption. The investigation suggests that the K- feature is less present in the atmosphere of HD209458b than in the one of HD189733b. This comparison confirms previous claims that the atmospheres of these two planets must have fundamentally different properties. | astro-ph.EP | astro-ph | MNRAS 000, 1 -- 5 (2019)
Preprint 13 September 2019
Compiled using MNRAS LATEX style file v3.0
The potassium absorption on HD189733b and HD209458b
Engin Keles,1 Matthias Mallonn,1 Carolina von Essen,2 Thorsten A. Carroll,1
Xanthippi Alexoudi,1 Lorenzo Pino,3 Ilya Ilyin,1 Katja Poppenhager,1
Daniel Kitzmann,4 Valerio Nascimbeni,5,6 Jake D. Turner,7,8 Klaus G. Strassmeier1
1Leibniz-Institut fur Astrophysik Potsdam (AIP), An der Sternwarte 16, 14482 Potsdam, Germany
2Stellar Astrophysics Centre, Department of Physics and Astronomy, Aarhus University, Ny Munkegade 120, 8000 Aarhus C, Denmark
3Leiden Observatory, Leiden University, Postbus 9513, 2300 RA Leiden, The Netherlands
4University of Bern, Center for Space and Habitability, Gesellschaftsstrasse 6, CH-3012, Bern, Switzerland
5Istituto Nazionale di Astrofisica, Osservatorio Astronomico di Padova, 35122 Padova, Italy
6Dipartimento di Fisica e Astronomia - Universta di Padova, Vicolo dell Osservatorio 3, I-35122 Padova
7Cornell University, Ithaca, New York, USA; 8University of Virginia, Charlottesville, Virginia, USA
Accepted XXX. Received YYY; in original form ZZZ
ABSTRACT
In this work, we investigate the potassium excess absorption around 7699A of the
exoplanets HD189733b and HD209458b. For this purpose, we used high spectral res-
olution transit observations acquired with the 2 × 8.4m Large Binocular Telescope
(LBT) and the Potsdam Echelle Polarimetric and Spectroscopic Instrument (PEPSI).
For a bandwidth of 0.8A, we present a detection > 7-σ with an absorption level of
0.18 % for HD189733b. Applying the same analysis to HD209458b, we can set 3-σ
upper limit of 0.09%, even though we do not detect a K- excess absorption. The inves-
tigation suggests that the K- feature is less present in the atmosphere of HD209458b
than in the one of HD189733b. This comparison confirms previous claims that the
atmospheres of these two planets must have fundamentally different properties.
Key words: exoplanet -- exoplanet atmosphere -- transmission spectroscopy -- star
1 INTRODUCTION
A suitable method for the characterization of planetary at-
mospheres is transmission spectroscopy (Seager & Sasselov
2000). During transit, a small fraction of the starlight is
absorbed or scattered by atoms and molecules, putting a
fingerprint on the stellar spectrum. One possibility to infer
these fingerprints is the "excess absorption" method, where
the flux of the spectral range of interest is integrated within a
bandwidth and divided by the flux within a reference band (a
spectral region where planetary absorption is not expected),
showing the absorption within the atmosphere during the
transit (Charbonneau et al. 2002). Especially hot giant plan-
ets are suitable targets of these kind of investigations due to
their large scale heights and short orbital periods, where
several atmospheric constituents have been successfully de-
tected e.g. sodium (Redfield et al. 2008; Snellen et al. 2008;
Casasayas-Barris et al. 2017), potassium (Sing et al. 2011),
titanium and iron (Hoeijmakers et al. 2018), hydrogen (Yan
& Henning 2018), and magnesium (Cauley et al. 2019).
For hot Jupiter atmospheres with T ∼ 1500K, the strongest
atomic absorber in the optical wavelengths are Na and K
(Fortney et al. 2010). Different investigations on HD189733b
© 2019 The Authors
and HD209458b have confirmed the presence of Na us-
ing low- and high resolution spectroscopy (see e.g.
for
HD209458b Charbonneau et al. (2002) and Snellen et al.
(2008) or for HD189733b Redfield et al. (2008) and Wyt-
tenbach et al. (2015)). However, the detection of K was not
yet assured from high resolution investigations for any exo-
planet, although attempts were made e.g. recently by Gibson
et al. (2019) investigating K on the exoplanet Wasp-31b.
Several investigations attempted to detect K on HD189733b
and HD209458b. For instance, Jensen et al. (2011) used the
Hobby-Eberly-Telescope and stated a non-detection of K for
both exoplanets. A tentative 2.5-σ detection of K in the at-
mosphere of HD189733b was claimed by Pont et al. (2013)
using the ACS camera at the Hubble Space Telescope. To
date, there is no significant detection of K for these two exo-
planets, neither in low- nor in high- resolution observations.
2 OBSERVATIONS
We observed one transit for HD189733b on October 11, 2017
at 01:47 -- 06:39 UT (PI: J.D. Turner, UVA) and one transit
for HD209458b on October 13, 2017 at 03:03 -- 08:04 UT
2
E. Keles et al.
Figure 1. Full PEPSI spectra (top) and K lines (bottom). Dashed lines mark the K lines (red) and control lines (colored) and the solid
blue line mark the telluric model for few telluric lines. Shaded area shows the planetary motion of around ±0.42A during the transit.
(PI: K.G. Strassmeier, AIP) using the PEPSI instrument
(Strassmeier et al. 2015, 2018b) with a 3.2 -- pixel resolution
of 130 000 at the LBT operated in the binocular mode. The
wavelength region of interest was covered with cross dis-
perser VI (7340 A -- 9070 A) in the polarimetric mode and
the beams recombined yielding the integral light spectrum.
The spectrograph is a white -- pupil fiber -- fed spectrograph
located in a pressure-controlled chamber at a constant pres-
sure, temperature and humidity to ensure constant refrac-
tive index of the air inside, providing radial velocity stability
about 5 m/s on the long term and less than 0.5 m/s on the
short term (Strassmeier et al. 2018a).
For the HD189733b transit, we obtained 24 spectra (15 out-
of-transit (OOT) and 9 in-transit (IT)) setting the exposure
time to 10 min. We exclude one OOT exposure due to a sys-
tematically lower relative flux determination. The resulting
signal-to-noise ratio (S/N) in the final processed data varied
due to increase in airmass during the night in the K line
at 7699 A from around 160 to 70 and at the continuum at
7700 A from around 540 to 270. For the HD209458b transit,
we obtained 17 spectra (9 OOT and 8 IT) setting the ex-
posure time to 10 min. The observations paused from 06:26
UT -- 07:19 UT due to bad weather conditions, leading to a
loss of phase coverage at the second part of the transit. The
resulting S/N in the final processed data varied due to an
increase in airmass during the night in the K line at 7699
A from around 230 to 80 and at the continuum at 7700 A
from around 425 to 160.
The image processing steps includes bias subtraction and
variance estimation of the source images, super -- master flat
field correction for the CCD spatial noise, echelle orders def-
inition from the tracing flats, scattered light subtraction,
wavelength solution for the ThAr images, optimal extrac-
tion of image slicers and cosmic spikes elimination of the
target image, wavelength calibration and merging slices in
each order, normalization to the master flat field spectrum
to remove CCD fringes and blaze function, a global 2D fit to
the continuum of the normalized image, and rectification of
all spectral orders in the image to a 1D spectrum. The soft-
ware numerical toolkit and graphical interface application
for PEPSI is described in Strassmeier et al. (2018a), and its
complete description is in preparation by Ilyin (2019).
The blaze function was removed by the division of the master
flat field spectrum. The residual spectra were fitted with a
low order 2D spline with subsequent rectification of spectral
orders where the overlapping parts of spectral orders were
averaged with their weights as the inverse variances of the
individual wavelength pixels. After that, the continuum of
individual spectra was corrected again with the mean spec-
trum as the weighted average of all the observed spectra.
The mean spectrum was normalized to its continuum and
each observed spectrum was divided by the mean normal-
ized spectrum. A low order spline fit to the ratio constitutes
the final continuum of the individual spectrum.
3 DATA ANALYSIS
The K doublet absorption lines are at 7698.98 A and 7664.92
A (see bottom panels of Figure 1). Since the latter line is
surrounded by a strong pair of telluric oxygen lines, we focus
our analysis only on the line at 7698.98 A. To demonstrate
the reliability of the excess absorption, we apply the same
procedure on several control lines, where we do not expect
planetary absorption (see e.g. Redfield et al. (2008)). Each
chosen control line is free of telluric contamination and has
a normalized flux level < 0.5 in the line centre. We use syn-
thetic stellar spectra to model the center-to-limb variation
(CLV) (see Section 3.3) to avoid a false-positive detection.
Top panels in Figure 1 show PEPSI spectra of the targets.
3.1 Telluric lines
Telluric lines are spectral features induced by Earths atmo-
sphere. To investigate if telluric line contamination has an
effect on the results (as the HD189733 spectra exhibit a weak
telluric line at 7699 A), each spectrum is telluric corrected
around the K lines (see bottom panel of Figure 1) and then
sampled to a common wavelength grid using a spline func-
tion. For this, we developed a telluric line corrector called
"Telluric Hapi Observation Reducer" (THOR), which uses
the "HITRAN Application Programming Interface" (HAPI)
as a basis (Kochanov et al. 2016). HAPI consists of differ-
ent python routines enabling the calculation of absorption
MNRAS 000, 1 -- 5 (2019)
Detection of potassium absorption on HD189733b
3
spectra using line-by-line data provided by the HITRAN
database (Gordon et al. 2017). THOR iterates HAPI for
different parameters (e.g. wavelength shifts, wavelength res-
olution, mean free path) until it reaches a χ2 -- min between
the observed and modelled telluric lines. To prove the relia-
bility of our telluric correction, we verify the increase in line
depth with airmass for the modelled telluric line at 7699 A
applying a Pearson correlation test, which result in a value of
0.90 (thus showing high correlation). As the telluric contam-
ination is weak, the excess level shown in Section 4 do not
change within 1-σ either applying telluric correction or not.
For further analysis, we use the telluric corrected spectra.
3.2 Excess absorption
For both targets, the planetary motion during the transit of
around ±16 km/s introduces a shift in the absorption wave-
length of approximately ±0.42A. To search for the K excess
absorption, we use bandwidths from 0.8 A to 10.0 A in steps
of 0.4 A (whereby one pixel corresponds to ∼0.016 mA). This
lower limit of the bandwidth ensures that a major part of
the planetary absorption (shifted by the planetary motion)
is inside the integration band. The lower limit also contains
a major part of the spectral line, avoiding artifacts intro-
duced by line shape changes (Snellen et al. 2008) due to the
Rossiter- McLaughlin- effect.
We use two methods to infer the planetary excess absorp-
tion. First, we follow the "traditional" way, integrating the
flux in the spectral range of interest and dividing it by the
mean of the integrated flux (at same bandwidth) at two ref-
erence positions adjacent to the red and blue. The center of
the reference bands for the HD189733b investigation are at
7693.40 A and 7700.40 A and for the HD209458b investiga-
tion at 7692.40 A and 7705.40 A. Second, we will infer the
excess absorption by using no reference bands at all (thus
only integrating the flux in the K-line). This is possible for
high S/N continuum normalized spectra with stable blaze
functions. For both of these methods, we made use of a 2-
order polynomial fit to the OOT data for normalization.
The error bars are calculated using the uncertainties sourc-
ing from the photon and readout noise propagated in the
data. They are scaled according to the standard deviation
of the OOT values, if the mean error was underestimated
compared to this standard deviation. We determine the ex-
cess absorption level using a Markov-Chain Monte Carlo
(MCMC) method provided by PyAstronomy (a collection
of Python routines implemented in the PyMC (Patil et al.
2010) and SciPy packages (Jones et al. 01 ), where we simu-
late a transit and scale the area of a planetary body to mimic
an absorbing atmosphere. We applied 100 000 MCMC itera-
tions (rejecting the first 30000 samples as burn-in phase)
to ensure that the final best-fit is provided. We checked
for convergence of the chains by splitting them in three
equally sized sub-groups and verified that their individual
mean agreed within 1-σ, whereby their individual uncer-
tainty agree within 3.5%. The uncertainties for the excess
absorption level correspond to a 1-σ confidence level.
3.3 Center-to-limb variation
As the planet covers different parts of the stellar surface dur-
ing its transit, the differential limb darkening between the
MNRAS 000, 1 -- 5 (2019)
line core and the stellar continuum leads to darkening or
brightening effects, which are evident in the excess absorp-
tion curves. A detailed discussion of the CLV effect and its
influence on excess absorption curves is shown e.g. in Czesla
et al. (2015) and Yan et al. (2017). We simulate a transit to
derive the CLV curves using synthetic stellar spectra, which
are calculated using the "spectrum" program by R.O. Gray
(Gray & Corbally 1994). For the calculation of the model at-
mospheres we used the Kurucz model (Kurucz 1970; Castelli
& Kurucz 2004). For HD189733b, we used an effective tem-
perature of 4875 K, a surface gravity of log g = 4.56 and
a metallicity of dex -0.03. For HD209458b, we used an ef-
fective temperature of 6092 K, a surface gravity of log g =
4.25 and a metallicity of dex 0.02 (see also Boyajian et al.
(2015) for comparison). We set µ = cos θ to specify the limb
angle, whereby µ = 0 refers to the limb and µ = 1 to the
center of the disk. We generated 20 spectra with limb angles
between µ = 0 and µ = 1 with a spacing of 0.05. To derive
the CLV curves, the stellar surface is mapped by a grid of
100 × 100 pixels containing the limb angle dependent fluxes
and the planetary surface is mapped by 31.6 × 31.6 pix-
els (whereby higher pixel values did not change significantly
the results, but increased the computational time). For each
planet position in front of the star, we calculate the total
stellar limb angle dependent flux at the position where we
expect planetary absorption and the reference band posi-
tion. The CLV- curve is then produced by the same way as
the excess curve. We validated our model by comparing the
CLV effect for the sodium excess absorption simulated for
HD189733b by Yan et al. (2017), getting similar results.
4 RESULTS
4.1 Investigating potassium on HD189733b
The left panel of Figure 2 shows the mean excess absorp-
tion curve for HD189733b at a 0.8 A integration band us-
ing adjacent reference bands (top) and no reference bands
(middle). The blue curve shows the MCMC fit and the green
solid line represents the CLV effect. As the CLV effect has
a negligible effect on the overall excess absorption, we ne-
glect it in the middle panel. The excess absorption levels
are 0.181 % ± 0.022 % (top) and 0.184 % ± 0.025 % (mid-
dle), thus similar within their error bars. This shows that
both methods are suitable to infer the planetary excess ab-
sorption. This absorption value corresponds to ∼13 scale
heights, hinting that the absorption must originate at high
altitudes in the atmosphere. The left bottom panel shows
the mean absorption for different bandwidths (using no ref-
erence bands) and the significance level. By increasing the
bandwidth more than 0.8 A, the absorption level decreases,
as expected, due to the integration of less atmospheric K ab-
sorption relative to the continuum flux. Also shown on the
right y-scale is the apparent planetary radius i.e. the radius
until which the atmosphere is opaque in units of the white
light radius. The significance level of the K detection for the
0.8 A bandwidth is determined with > 7-σ with respect to
the zero level. For the remaining other bandwidths, the de-
termined excess absorption levels are also above 3-σ with
respect to the zero level.
Comparing our results to other investigations, we can qual-
itatively confirm the tentative K- detection of Pont et al.
4
E. Keles et al.
BW[A] EA[%]
0.8 0.184 ± 0.022
1.2 0.139 ± 0.017
1.6 0.116 ± 0.016
2.0 0.108 ± 0.021
2.4 0.087 ± 0.020
2.8 0.084 ± 0.018
3.2 0.079 ± 0.018
3.6 0.074 ± 0.018
4.0 0.068 ± 0.018
4.4 0.064 ± 0.017
4.8 0.060 ± 0.017
5.2 0.057 ± 0.016
5.6 0.060 ± 0.015
6.0 0.058 ± 0.014
6.4 0.059 ± 0.014
6.8 0.054 ± 0.014
7.2 0.049 ± 0.013
7.6 0.049 ± 0.013
8.0 0.050 ± 0.012
BW[A] EA[%]
0.8 -0.047 ± 0.028
1.2 -0.011 ± 0.021
1.6 -0.009 ± 0.018
2.0 -0.004 ± 0.016
2.4 -0.014 ± 0.018
2.8 -0.013 ± 0.017
3.2 -0.019 ± 0.017
3.6 -0.006 ± 0.017
4.0 -0.005 ± 0.016
4.4 0.016 ± 0.017
4.8 0.025 ± 0.018
5.2 0.017 ± 0.019
5.6 0.013 ± 0.018
6.0 0.013 ± 0.019
6.4 0.018 ± 0.019
6.8 0.031 ± 0.020
7.2 0.028 ± 0.021
7.6 0.030 ± 0.021
8.0 0.028 ± 0.021
Figure 2. K excess absorption (EA) for HD189733b (left) and HD209458b (right). EA curves at 0.8 A bandwidth (BW) using reference
bands (top) and no reference bands (middle). Bottom panel shows the EA level for different BW. Green line shows the CLV-curve and
the blue line the planetary absorption (both mirrored to mimic the second half of the transit for HD209458b). Tables show EA up to 8A.
(2013). Other way around, Jensen et al. (2011) investigated
at a resolution of 60 000 potassium on HD189733b, stating
a non-detection. In their work, they averaged the IT and
OOT spectra from several observations to produce one mas-
ter Fin/Fout spectrum and determined thereafter from this
the excess absorption. To do this, they subtracted the inte-
grated flux from the line core at the K-line from the mean
of two reference band fluxes using an 8A integration band
deriving an excess level of (-0.77 ± 1.04) x 10-4. Determining
the excess absorption at 8A bandwidth, we infer an excess
absorption value of (4.95 ± 1.20) x 10-4, deviating signifi-
cantly more than 3-σ from their findings. Possible reasons
which could explain this discrepancy are e.g. telluric lines
and stellar activity in some of the observations which can
affect the combined master spectra or even the difference in
the technique used to derive the excess level.
Figure 3 shows the excess level at a 0.8 A integration band
(using no reference bands) for several control lines which all
lie within 3-σ around zero opposite to the excess level at the
K- line at 7698.98 A (red square). The significance level of
the K absorption remains > 3-σ with respect to the standard
deviation of the excess level for the control lines, presenting
a strong evidence on the atmospheric K- absorption.
4.2 Stellar activity of HD189733
As the star HD189733 is an active K-dwarf, the stellar
variability could cause errors in the determined excess lev-
els e.g. by flares (Klocov´a et al. 2017; Khalafinejad et al.
2017; Cauley et al. 2018). Cauley et al. (2018) showed for
HD189733b that the transit of active latitudes with bright
facular and plage regions can cause emission feature in stel-
lar lines. Also the strong magnetic field of HD189733 could
have a significant effect on the emission in lines (Cauley et al.
2017). Possibly related to the activity, we see an emission-
like feature in several stellar line cores as e.g. in Na i (8183.28
A), Fe i (7511.03 A) and the investigated K i line (7698.98
A). As the integrated flux in the line core is low compared
to the integrated flux over the line width (of at least 0.8A),
this has a negligible influence on the result. Moreover, this
feature increases over the night and it is not restricted to the
transit, opposite to the excess absorption in Figure 2, which
appears only during the transit, making us confident about
the K absorption within the atmosphere of HD189733b.
4.3 Investigating potassium on HD209458b
The right panel in Figure 2 shows the same approach on
HD209458b as for HD189733b. Similar to HD189733b, the
CLV- effect is very weak and does not affect the result. In
contrast, there is no excess absorption evident at the in-
vestigated bandwidths, either using reference bands or not.
Moreover, the result suggests an emission like behaviour at
low bandwidths. Assuming zero excess absorption to not un-
derestimate the error (as a negative excess level would sug-
gest emission either than absorption), we determine a 3-σ
upper limit of around 0.084% at a bandwidth of 0.8A. Con-
cluding, in comparison to HD189733b, HD209458b shows no
significant absorption of potassium in its atmosphere.
5 DISCUSSION
Although both targets could have experienced different evo-
lutionary scenarios, changing their primordial atmospheres
leading to different atmospheric composition and proper-
ties, we make the assumption that the initial alkali metal
abundances could be similar for both targets, as they or-
bit a host star of solar metallicity (Boyajian et al. 2015).
As gas giants form and accrete H/He-dominated gas, they
also accrete planetesimals that enrich their envelope in met-
als, thus they can possess even a higher absolute metellicity
compared to the parent star (Nikolov et al. 2018). Indepen-
dent of any formation location in the protoplanetary disc,
the alkali metal ratios should not be affected significantly,
as the accretion of planetesimals will enrich both (here Na
MNRAS 000, 1 -- 5 (2019)
Detection of potassium absorption on HD189733b
5
REFERENCES
Barman T., 2007, ApJ, 661, L191
Boyajian T., et al., 2015, MNRAS, 447, 846
Casasayas-Barris N., Palle E., Nowak G., Yan F., Nortmann L.,
Murgas F., 2017, A&A, 608, A135
Castelli F., Kurucz R. L., 2004, arXiv Astrophysics e-prints,
Cauley P. W., Redfield S., Jensen A. G., 2017, AJ, 153, 217
Cauley P. W., Kuckein C., Redfield S., Shkolnik E. L., Denker C.,
Llama J., Verma M., 2018, AJ, 156, 189
Cauley P. W., Shkolnik E. L., Ilyin I., Strassmeier K. G., Redfield
S., Jensen A., 2019, AJ, 157, 69
Charbonneau D., Brown T. M., Noyes R. W., Gilliland R. L.,
2002, ApJ, 568, 377
Czesla S., Klocov´a T., Khalafinejad S., Wolter U., Schmitt
J. H. M. M., 2015, A&A, 582, A51
Fortney J. J., Sudarsky D., Hubeny I., Cooper C. S., Hubbard
W. B., Burrows A., Lunine J. I., 2003, ApJ, 589, 615
Fortney J. J., Shabram M., Showman A. P., Lian Y., Freedman
R. S., Marley M. S., Lewis N. K., 2010, ApJ, 709, 1396
Gibson N. P., de Mooij E. J. W., Evans T. M., Merritt S., Nikolov
N., Sing D. K., Watson C., 2019, MNRAS, 482, 606
Gordon I. E., et al., 2017, J. Quant. Spectrosc. Radiative Transfer,
203, 3
Gray R. O., Corbally C. J., 1994, AJ, 107, 742
Helling C., et al., 2016, MNRAS, 460, 855
Hoeijmakers H. J., et al., 2018, Nature, 560, 453
Jensen A. G., Redfield S., Endl M., Cochran W. D., Koesterke
L., Barman T. S., 2011, ApJ, 743, 203
Jones E., Oliphant T., Peterson P., et al., 2001 -- , SciPy: Open
source scientific tools for Python, http://www.scipy.org/
Khalafinejad S., et al., 2017, A&A, 598, A131
Klocov´a T., Czesla S., Khalafinejad S., Wolter U., Schmitt
J. H. M. M., 2017, A&A, 607, A66
Kochanov R. V., Gordon I. E., Rothman L. S., Wcis(cid:32)lo P., Hill C.,
Wilzewski J. S., 2016, J. Quant. Spectrosc. Radiative Trans-
fer, 177, 15
Kurucz R. L., 1970, SAO Special Report, 309
Lines S., et al., 2018, A&A, 615, A97
Lodders K., 1999, ApJ, 519, 793
Nikolov N., et al., 2018, Nature, 557, 526
Patil A., Huard D., Fonnesbeck C., 2010, Journal of Statistical
Software, Articles, 35, 1
Pont F., Sing D. K., Gibson N. P., Aigrain S., Henry G., Husnoo
N., 2013, MNRAS, 432, 2917
Redfield S., Endl M., Cochran W. D., Koesterke L., 2008, ApJ,
673, L87
Seager S., Sasselov D. D., 2000, ApJ, 537, 916
Sing D. K., Vidal-Madjar A., D´esert J.-M., Lecavelier des Etangs
A., Ballester G., 2008a, ApJ, 686, 658
Sing D. K., Vidal-Madjar A., Lecavelier des Etangs A., D´esert
J.-M., Ballester G., Ehrenreich D., 2008b, ApJ, 686, 667
Sing D. K., et al., 2011, A&A, 527, A73
Snellen I. A. G., Albrecht S., de Mooij E. J. W., Le Poole R. S.,
2008, A&A, 487, 357
Strassmeier K. G., et al., 2015, Astronomische Nachrichten, 336,
324
Strassmeier K. G., Ilyin I., Steffen M., 2018a, A&A, 612, A44
Strassmeier K. G., Ilyin I., Weber M., 2018b, A&A, 612, A45
Vidal-Madjar A., et al., 2011, A&A, 527, A110
Wyttenbach A., Ehrenreich D., Lovis C., Udry S., Pepe F., 2015,
A&A, 577, A62
Yan F., Henning T., 2018, Nature Astronomy, 2, 714
Yan F., Pall´e E., Fosbury R. A. E., Petr-Gotzens M. G., Henning
T., 2017, A&A, 603, A73
This paper has been typeset from a TEX/LATEX file prepared by
the author.
Figure 3. Mean excess absorption for several control lines com-
pared to K (red square) at 0.8 A bandwidth for HD189733b.
and K) similarly. Therefore, to make an assumption about
the expected K excess level on HD209458b, we compare the
previously detected Na excess level on both planets assum-
ing similar Na/K -- ratio. Although the Na excess absorp-
tion level for HD189733b is around two times larger than
for HD2093458 (Snellen et al. 2008; Jensen et al. 2011), it
corresponds to a similar variation in apparent planet ra-
dius (Snellen et al. 2008). Translating the K excess absorp-
tion of 0.184% on HD189733b to the variation in apparent
planet radius, one would expect a K absorption of around
0.108% for HD209458b. This value is larger than our upper
3-σ limit of 0.084%, indicating that K could be depleted on
HD209458b, either by condensation processes in the lower
and/or photo-ionization processes in the upper atmosphere.
Such an indication is also suggested by Sing et al. (2008b,a),
who observed a broad absorption plateau at lower altitudes
and a narrow absorption line core for the Na feature. The
authors argued that condensation on the night side of the
planet can lead to the loss of atmospheric Na, as the atmo-
spheric temperatures become cool enough to condensate Na
and lead to the absence on high altitudes. This was also con-
firmed by Vidal-Madjar et al. (2011) showing temperature-
pressure-profile simulations to match the observations by
Sing et al. (2008b,a) and Snellen et al. (2008).
Comparing to those findings, the interpretation of the K de-
tection is puzzling. The condensation of Na to Na2S as well
as NaCl happens at higher temperature than the condensa-
tion of K to KCl (Lodders 1999), so that one would expect K
to be more abundant than Na, in contrast with what is ob-
served. An alternative depletion process is photo-ionization.
Potassium has a slightly lower ionization energy than Na,
leading to easier photo-ionization of K (Fortney et al. 2003;
Barman 2007). Then, one may expect that depletion of K is
stronger on HD189733b (which orbits an active star) than on
HD209458b, not being in agreement with our findings, mak-
ing the atmospheric conditions on HD209458b puzzling.
Although both planets have similar properties (e.g. size,
orbital period and equilibrium temperature) there are de-
viations especially on their cloud properties (Lines et al.
2018). Modelling approaches suggest that cloud formation
on HD189733b originates at lower pressure levels due to its
higher gravity and cloud particle density (Lines et al. 2018)
opposite to dust clouds on HD209458b, which extent to large
areas on the atmosphere (Helling et al. 2016).
These differences (among others) could lead to different con-
densation chemistry on both targets, affecting the alkali de-
pletion processes and thus the implications discussed above.
MNRAS 000, 1 -- 5 (2019)
|
1006.0114 | 1 | 1006 | 2010-06-01T11:38:26 | First results of Herschel/PACS observations of Neptune | [
"astro-ph.EP"
] | We report on the initial analysis of a Herschel/PACS full range spectrum of Neptune, covering the 51-220 micrometer range with a mean resolving power of ~ 3000, and complemented by a dedicated observation of CH4 at 120 micrometers. Numerous spectral features due to HD (R(0) and R(1)), H2O, CH4, and CO are present, but so far no new species have been found. Our results indicate that (i) Neptune's mean thermal profile is warmer by ~ 3 K than inferred from the Voyager radio-occultation; (ii) the D/H mixing ratio is (4.5+/-1) X 10**-5, confirming the enrichment of Neptune in deuterium over the protosolar value (~ 2.1 X 10**-5); (iii) the CH4 mixing ratio in the mid stratosphere is (1.5+/-0.2) X 10**-3, and CH4 appears to decrease in the lower stratosphere at a rate consistent with local saturation, in agreement with the scenario of CH4 stratospheric injection from Neptune's warm south polar region; (iv) the H2O stratospheric column is (2.1+/-0.5) X 10**14 cm-2 but its vertical distribution is still to be determined, so the H2O external flux remains uncertain by over an order of magnitude; and (v) the CO stratospheric abundance is about twice the tropospheric value, confirming the dual origin of CO suspected from ground-based millimeter/submillimeter observations. | astro-ph.EP | astro-ph | Astronomy&Astrophysicsmanuscript no. 14600
June 3, 2018
c(cid:13) ESO 2018
Letter to the Editor
First results of Herschel/ PACS observations of Neptune ⋆
E. Lellouch1, P. Hartogh2, H. Feuchtgruber3, B. Vandenbussche4, T. de Graauw5, R. Moreno1, C. Jarchow2,
T. Cavali´e2, G. Orton6, M. Banaszkiewicz7, M.I. Blecka7, D. Bockel´ee-Morvan1, J. Crovisier1, T. Encrenaz1,
T. Fulton8, M. Kuppers9, L.M. Lara10, D.C. Lis11, A.S. Medvedev2, M. Rengel2, H. Sagawa2, B. Swinyard12,
S. Szutowicz7, F. Bensch13, E. Bergin14, F. Billebaud15, N. Biver1, G.A. Blake6, J.A.D.L. Blommaert4, J. Cernicharo16,
R. Courtin1, G.R. Davis17, L. Decin4, P. Encrenaz18, A. Gonzalez2, E. Jehin19, M. Kidger20, D. Naylor21, G.
Portyankina22, R. Schieder23, S. Sidher12, N. Thomas22, M. de Val–Borro2, E. Verdugo20, C. Waelkens4, H. Walker12,
H. Aarts5, C. Comito24, J.H. Kawamura6, A. Maestrini18, T. Peacocke25, R. Teipen23, T. Tils23, and K. Wildeman5
0
1
0
2
n
u
J
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
1
1
0
.
6
0
0
1
:
v
i
X
r
a
(Affiliations can be found after the references)
Draft June 3, 2018; received ???; accepted ???
ABSTRACT
We report on the initial analysis of a Herschel / PACS full range spectrum of Neptune, covering the 51-220 µm range with a mean resolving power
of ∼3000, and complemented by a dedicated observation of CH4 at 120 µm. Numerous spectral features due to HD (R(0) and R(1)), H2O, CH4,
and CO are present, but so far no new species have been found. Our results indicate that (i) Neptune's mean thermal profile is warmer by ∼3 K than
inferred from the Voyager radio-occultation; (ii) the D/H mixing ratio is (4.5±1)×10−5, confirming the enrichment of Neptune in deuterium over
the protosolar value (∼2.1×10−5); (iii) the CH4 mixing ratio in the mid stratosphere is (1.5±0.2)×10−3, and CH4 appears to decrease in the lower
stratosphere at a rate consistent with local saturation, in agreement with the scenario of CH4 stratospheric injection from Neptune's warm south
polar region; (iv) the H2O stratospheric column is (2.1±0.5)×1014 cm−2 but its vertical distribution is still to be determined, so the H2O external
flux remains uncertain by over an order of magnitude; and (v) the CO stratospheric abundance is about twice the tropospheric value, confirming
the dual origin of CO suspected from ground-based millimeter/submillimeter observations.
Key words. Planets and satellites: individual: Neptune; Techniques: spectroscopic; Infrared: solar system; Radio lines: solar system)
1. Introduction
Table 1. Summary of observations
Neptune's thermal emission has been initally explored from the
ground in the 8-13 µm window and in the millimeter range and
by the Voyager spacecraft in 1989, but detailed views of its
spectrum had to await sensitive instrumentation onboard ISO
(see review in B´ezard et al. 1999a), Spitzer (Meadows et al.
2008) and recently AKARI (Fletcher et al. 2010). Altogether,
these observations have revealed a surprisingly rich composi-
tion of Neptune's stratosphere, including numerous hydrocar-
bons (CH4, C2H2, C2H6, CH3, C2H4, CH3C2H, C4H2), oxygen-
bearing species (CO, CO2, and H2O), HCN, as well as deu-
terium species CH3D and HD. Favorable factors for observing
minor species in Neptune's atmosphere are (i) its relatively warm
stratosphere (∼140 K at 1 mbar) that enhances IR emission; and
(ii) Neptune's large internal heat source that results in rapid con-
vection updrafting minor disequilibrium species, notably CO,
up to observable levels. Neptune's submillimeter spectrum long-
wards of 50 µm has been observed by ISO/LWS (Burgdorf et
al. 2003), but the signal–to–noise ratio in the data was not high
enough to reveal spectral features. In this paper, we report the
first results from observations of Neptune at 51-220 µm (195–
45 cm−1) with the PACS instrument onboard Herschel (Pilbratt
et al. 2010), performed in the framework of the KP-GT "Water
and Related Chemistry in the Solar System", also known as
"Herschel Solar System observations" (Hartogh et al. 2009).
⋆ Herschel is an ESA space observatory with science instruments
provided by European-led Principal Investigator consortia and with im-
portant participation from NASA.
Obs. ID
Start Time
[UTC]
Tobs
[min.]
Rangea
[µm]
1342186536
1342186537
1342186538
1342186539
1342186540
1342186571
30-Oct-2009 00:58:36
30-Oct-2009 03:01:48
30-Oct-2009 05:22:32
30-Oct-2009 08:53:20
30-Oct-2009 11:31:41
31-Oct-2009 14:35:00
116
133
203
151
236
82
51-72k, 102-145n
51-62l, 150-186n
60-73l, 180-220n
68-85m, 120-171n
82-102m, 165-220n
118.4-120.9n
a grating order and filter: k = 2A, l = 3A, m = 2B, n = 1 red
2. Herschel/ PACS observations
All observations (Table 1) were carried out in chopped-nodded
PACS range spectroscopy modes (Poglitsch et al. 2010) at high
spectral sampling density. The entire spectral range of PACS
has been measured at full instrumental resolution λ/δλ rang-
ing from 950 to 5500 depending on wavelength and grating or-
der (PACS Observers Manual 2010). A summary of the observa-
tions is given in Table 1. Since blue and red spectrometer data
are acquired in parallel, several spectral ranges have been ob-
served in overlap. Given the instrumental spatial pixel size of
9.4"×9.4", Neptune (2.297" as seen from Herschel) can be con-
sidered as a point source, and the analyzed spectra therefore
2
Lellouch et al.: PACS observations of Neptune
m
m
m
m
m
m
m
m
m
m
m
m
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
n
n
n
n
n
n
n
n
n
n
n
n
i
i
i
i
i
i
i
i
i
i
i
i
t
t
t
t
t
t
t
t
t
t
t
t
n
n
n
n
n
n
n
n
n
n
n
n
o
o
o
o
o
o
o
o
o
o
o
o
c
c
c
c
c
c
c
c
c
c
c
c
/
/
/
/
/
/
/
/
/
/
/
/
e
e
e
e
e
e
e
e
e
e
e
e
n
n
n
n
n
n
n
n
n
n
n
n
L
L
L
L
L
L
L
L
L
L
L
L
i
i
i
i
i
i
i
i
i
i
i
i
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.15
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1.05
1
1
1
1
1
1
1
1
1
1
1
1
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.95
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.85
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.8
Observed
Observed
Observed
Observed
Observed
Observed
Observed
Observed
Observed
Observed
Observed
Observed
CH4
CH4
CH4
CH4
H2O
H2O
H2O
CO
CO
HD
60
60
60
60
60
60
60
60
60
60
60
60
80
80
80
80
80
80
80
80
80
80
80
80
100
100
100
100
100
100
100
100
100
100
100
100
120
120
120
120
120
120
120
120
120
120
120
120
140
140
140
140
140
140
140
140
140
140
140
140
160
160
160
160
160
160
160
160
160
160
160
160
180
180
180
180
180
180
180
180
180
180
180
180
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Fig. 1. Composite PACS spectrum of Neptune, expressed in line/local continuum ratios. For spectral ranges covered more than once
(Table 1), the observation with the highest resolution has been selected. The region beyond 190 µm is not shown, owing to severe
mixing of spectrometer orders. The bottom curves are synthetic spectra at appropriate spectral resolution that show the contributions
of CH4, H2O, CO, and HD lines.
originate only in the central spatial pixel of the integral field
spectrometer.
3. Analysis and discussion
3.1. ThermalprofileandD/Habundance
Starting from Level 0 products, the processing of all ob-
servations was carried by standard PACS pipeline modules
(Poglitsch et al. 2010) up to Level 1. Individual spectral pixels
were then scaled onto their common mean in order to improve
the removal of signal outliers caused by cosmic ray hits. After
application of an iterative σ-clipping, adapted to the instrumen-
tal resolution, the remaining data were rebinned onto an over-
sampled wavelength grid to ensure conservation of spectral res-
olution. The absolute flux calibration of the instrument and im-
provements on the relative spectral response function are still in
progress. Therefore the resulting spectrum was then divided by
its continuum, to be robust against forthcoming calibration up-
dates. The composite spectrum is shown in Fig.1. It shows emis-
sion signatures due to CH4, H2O, CO, as well as the R(0) and
R(1) lines of HD at 112 and 56 µm, seen respectively in absorp-
tion and emission. At this stage of the data reduction, features
below ∼0.5-1 % contrast must be treated with caution. No new
species are detected at this level.
A dedicated line–scan high S/N observation of the CH4
119.6 µm rotational line was also acquired in order to get a high
precision measure of the CH4 stratospheric abundance.
Observations were analyzed by means of a standard radiative
transfer code, in which the outgoing radiance from Neptune was
integrated over all emission angles. The effective spectral reso-
lution as a function of wavelength was determined by fitting the
widths of the H2O lines, whose profile is purely instrumental. We
initially considered thermal profiles inferred in previous work
(Marten et al. 2005, B´ezard et al. 1998, Fletcher et al. 2010, re-
spectively from ground-based, ISO, and AKARI observations).
Below about 0.5 bar, all of them follow the Voyager radio-
occultation profile (Lindal 1992, see also Moses et al. 2005).
Above this level, these profiles diverge significantly, showing ex-
cursions of ∼5 K over 10-200 mbar, and even larger dispersion
(∼10-20 K) at lower pressures. Over 50-200 µm, Neptune's con-
tinuum is formed near the 500 mbar level (TB ∼ 59 K). The HD
lines typically probe the 10-500 mbar range (peak contribution
near 2 mbar at line center). Because they show a contrasted ab-
sorption/emission appearance and because HD is vertically well
mixed, they provide a sensitive thermometer in this region. For
HD, we used the same linestrengths as in Feuchtgruber et al.
(1999). We found that the Fletcher et al. (2010) nominal pro-
file (their Fig. 5) allowed a much better fit of the HD lines than
the other two profiles, and achieved optimum fit for tempera-
tures equal to 0.9×Fletcher + 0.1×Marten (Fig. 2). This gives
54.5 K at the tropopause, ∼3 K higher than in Lindal (1992).
Lellouch et al.: PACS observations of Neptune
3
m
m
m
m
m
m
m
m
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
u
n
n
n
n
n
n
n
n
i
i
i
i
i
i
i
i
t
t
t
t
t
t
t
t
n
n
n
n
n
n
n
n
o
o
o
o
o
o
o
o
c
c
c
c
c
c
c
c
/
/
/
/
/
/
/
/
e
e
e
e
e
e
e
e
n
n
n
n
n
n
n
n
L
L
L
L
L
L
L
L
i
i
i
i
i
i
i
i
m
m
m
m
m
u
u
u
u
u
u
u
u
u
u
n
n
n
n
n
i
i
i
i
i
t
t
t
t
t
n
n
n
n
n
o
o
o
o
o
c
c
c
c
c
/
/
/
/
/
e
e
e
e
e
n
n
n
n
n
L
L
L
L
L
i
i
i
i
i
1.06
1.06
1.06
1.06
1.06
1.06
1.06
1.06
1.04
1.04
1.04
1.04
1.04
1.04
1.04
1.04
1.02
1.02
1.02
1.02
1.02
1.02
1.02
1.02
1
1
1
1
1
1
1
1
0.98
0.98
0.98
0.98
0.98
0.98
0.98
0.98
56
56
56
56
56
56
56
56
56.1
56.1
56.1
56.1
56.1
56.1
56.1
56.1
56.2
56.2
56.2
56.2
56.2
56.2
56.2
56.2
56.3
56.3
56.3
56.3
56.3
56.3
56.3
56.3
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
56.4
56.4
56.4
56.4
56.4
56.4
56.4
56.4
56.5
56.5
56.5
56.5
56.5
56.5
56.5
56.5
1.02
1.02
1.02
1.02
1.02
1.01
1.01
1.01
1.01
1.01
1
1
1
1
1
0.99
0.99
0.99
0.99
0.99
0.98
0.98
0.98
0.98
0.98
0.97
0.97
0.97
0.97
0.97
0.96
0.96
0.96
0.96
0.96
0.95
0.95
0.95
0.95
0.95
0.94
0.94
0.94
0.94
0.94
111
111
111
111
111
111.5
111.5
111.5
111.5
111.5
112
112
112
112
112
112.5
112.5
112.5
112.5
112.5
113
113
113
113
113
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Fig. 2. Neptune's temperature and abundance profiles. CH4 pro-
files condensing (thick red line) or not (thin red line) in the
stratosphere are considered. For H2O, profiles A and B are those
of Feuchtgruber et al. (1997), multiplied by 0.95 and 0.9, respec-
tively, and the "uniform" profile has a mixing ratio of 0.85 ppb
above the condensation level. For CO, the profiles of Lellouch
et al. (2005) and Fletcher et al. (2010) are shown. The black line
shows the inferred temperature profile.
Fig. 3. Neptune's spectrum in the 56.0–56.5 and 111–113 µm
ranges, showing HD lines at 56.25 µm (R(1)) and 112.1 µm
(R(0)). Thick red line: model for HD/H2 = 9×10−5 and the nom-
inal thermal profile of Fig. 2. Thin pink lines: same for HD/H2
= 6×10−5 and 12×10−5. Green: model for HD/H2 = 9×10−5 and
Marten's et al. (2005) thermal profile). Note also the water line
at 56.35 µm, well fitted by profile A in Fig. 2. The upper and
lower blue lines show models for this H2O profile multiplied
and divided by 1.5.
Given Neptune's temperature field as inferred from Voyager
measurements (Conrath et al. 1998), this is probably related
to the high latitude (42◦S) of the Voyager occultations. Based
on mid–infrared measurements of ethane, Hammel et al. (2006)
also found enhanced temperatures (but at sub-mbar levels) com-
pared to Voyager, a likely consequence of seasonal variability.
Although the HD lines do not constrain temperatures above the
1 mbar level (needed in particular for analyzing the H2O lines),
we retained the 0.9×Fletcher + 0.1×Marten combination for all
levels. We determined HD/H2 = (9±2)×10−5, i.e. a D/H ratio
of (4.5±1)×10−5 (Fig. 3). This is nominally less than but consis-
tent with the (6.5+2.5
−1.5)×10−5 value inferred by Feuchtgruber et al.
(1999) from observations of the R(2) line of HD by ISO/SWS,
and confirms that Neptune is enriched in deuterium compared
to the protosolar value (∼2.1×10−5) represented by Jupiter and
Saturn (Lellouch et al. 2001). We defer a joint analysis of ISO
and Herschel data to future work.
m
m
m
m
m
u
u
u
u
u
u
u
u
u
u
n
n
n
n
n
i
i
i
i
i
t
t
t
t
t
n
n
n
n
n
o
o
o
o
o
c
c
c
c
c
/
/
/
/
/
e
e
e
e
e
n
n
n
n
n
L
L
L
L
L
i
i
i
i
i
m
m
m
u
u
u
u
u
u
n
n
n
i
i
i
t
t
t
n
n
n
o
o
o
c
c
c
/
/
/
e
e
e
n
n
n
L
L
L
i
i
i
1.06
1.06
1.06
1.06
1.06
1.05
1.05
1.05
1.05
1.05
1.04
1.04
1.04
1.04
1.04
1.03
1.03
1.03
1.03
1.03
1.02
1.02
1.02
1.02
1.02
1.01
1.01
1.01
1.01
1.01
1
1
1
1
1
0.99
0.99
0.99
0.99
0.99
1.03
1.03
1.03
1.02
1.02
1.02
1.01
1.01
1.01
1
1
1
0.99
0.99
0.99
0.98
0.98
0.98
Observed
Observed
Observed
Observed
Observed
119.2
119.2
119.2
119.2
119.2
119.4
119.4
119.4
119.4
119.4
119.6
119.6
119.6
119.6
119.6
119.8
119.8
119.8
119.8
119.8
120
120
120
120
120
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Observed
Observed
Observed
158.6
158.6
158.6
159
159
159
159.4
159.4
159.4
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
159.8
159.8
159.8
3.2. Methane,water,andcarbonmonoxideabundancesand
profiles
Methane has been observed in Neptune's stratosphere with
a range of abundances exceeding the saturation value at the
tropopause cold trap (e.g. Baines and Hammel, 1994). The PACS
spectrum shows several rotational lines of CH4 in emission over
80–160 µm. Thanks to the mild temperature dependence of the
Planck function in this spectral range, these lines are well suited
to determination of the CH4 stratospheric abundance. We as-
sumed a CH4 abundance of 2 % in the deep troposphere, then
following the saturation law. In the stratosphere, the CH4 pro-
file was characterized by its high–altitude mixing ratio (qCH4)
and assumed to follow local saturation below the condensation
point near 40 mbar. Utilizing the Boudon et al. (2010) results on
the absolute CH4 line strengths and in particular using the high
S/N dedicated CH4 120 µm line scan (Fig. 4), we determined
qCH4 = (1.5±0.2)×10−3, consistent with B´ezard et al. (1999b)
((0.5–2)×10−3) but only marginally with Fletcher et al. (2010)
Fig. 4. Methane lines at 119.6 µm (Obs.ID 1342186571) and
159.3 µm (from Obs.ID 1342186537). Red: Model for strato-
spheric qCH4 = 0.0015 above the stratospheric saturation level
(thick red line in Fig. 2). Green curves: same, but for qCH4 =
0.0020 (upper curve) and 0.0010 (lower curve). Blue: Model in
which qCH4 = 0.0025 down to ∼800 mbar (thin red line in Fig.
2).
((0.9±0.3)×10−3). Because of the progressive increase of the
continuum level longwards of 100 µm, the CH4 features at 137
µm and particularly 159 µm are sensitive to the CH4 amount in
the lower stratosphere. An alternate assumption would be that
the CH4 is supersaturated there, as could perhaps result from
strong convective overshoot. This situation leads, however, to
unobserved absorption wings at 159 µm and to inconsistent mix-
ing ratios for the different lines (Fig. 4). A 1.5×10−3 mixing ra-
tio is ∼10 times greater than allowed by the 56 K cold trap, and
consistent with saturation at 60 K. The most probable origin of
this elevated stratospheric abundance is that CH4 leaks from the
4
Lellouch et al.: PACS observations of Neptune
hot (62–66 K at the tropopause) Southern region (Orton et al.
2007) and is redistributed planetwide by global circulation. A
combined analysis of the PACS, ISO, Spitzer, and AKARI data
in terms of stratospheric methane and temperature profile will be
performed in the future.
The presence of H2O in giant planet stratospheres, includ-
ing Neptune's, was established from ISO/SWS 30–45 µm spec-
tra (Feuchtgruber et al. 1997), demonstrating the existence of an
external oxygen supply. In Neptune's case, ISO observations de-
termined a (2–4)×1014 cm−2 column density, but did not estab-
lish the water vertical profile, a parameter needed to derive the
rate at which water is removed by vertical mixing and condensa-
tion and to infer the input flux of water. More than 20 H2O lines,
encompassing over a range in opacity of more than an order of
magnitude (∼0.2 to 2.5), are detected in the PACS spectrum. If
uniformly mixed above the condensation level near 1.2 mbar,
the water mixing ratio is qH2O = (0.85±0.2) ppb, and its col-
umn density is (2.1±0.5)×1014 cm−2. Following Feuchtgruber
et al. (1997), we also considered H2O vertical profiles resulting
from transport models, characterized by the eddy diffusion co-
efficient profile (profiles "A" and "B", see Fig. 2). For a given
vertical profile, the water amounts we determined from the data
were identical, to within 10 %, to the values inferred from ISO.
However, the associated external fluxes vary strongly (1.4×105
cm−2s−1 for model A and 9×106 cm−2s−1 for model B). We leave
the detailed retrieval of Neptune's water profile (including PACS
targeted observations of several weak lines and a deep 557 GHz
HIFI observation) for the future. For the time being, an elemen-
tary analysis based on the integrated linewidths favors profile A
over the other two water profiles (Fig.5), suggesting that the wa-
ter mixing ratio increases with altitude over 0.1–1 mbar.
more than a factor of 2 (1×10−6 and 2.2×10−6, respectively) on
the stratospheric CO abundance (Fig. 2). Support for the Hesman
et al. value was reported from the detection of CO fluorescence
at 4.7 µm by AKARI (Fletcher et al. 2010), from which a 2.5
ppm abundance of CO above the 10-mbar pressure level was in-
ferred. We find here that the CO lines longward of 150 µm (Fig.
6) instead imply a CO stratospheric abundance of ∼1 ppm, in
agreement with Lellouch et al. (2005). The detailed determina-
tion of the CO profile will be possible from combined analysis
of PACS, SPIRE, and new broadband ground-based millimeter
data.
m
m
m
m
m
m
u
u
u
u
u
u
u
u
u
u
u
u
n
n
n
n
n
n
i
i
i
i
i
i
t
t
t
t
t
t
n
n
n
n
n
n
o
o
o
o
o
o
c
c
c
c
c
c
/
/
/
/
/
/
e
e
e
e
e
e
n
n
n
n
n
n
L
L
L
L
L
L
i
i
i
i
i
i
1.06
1.06
1.06
1.06
1.06
1.06
1.04
1.04
1.04
1.04
1.04
1.04
1.02
1.02
1.02
1.02
1.02
1.02
1
1
1
1
1
1
0.98
0.98
0.98
0.98
0.98
0.98
0.96
0.96
0.96
0.96
0.96
0.96
0.94
0.94
0.94
0.94
0.94
0.94
0.92
0.92
0.92
0.92
0.92
0.92
Observed
Observed
Observed
Observed
Observed
Observed
CO Fletcher
CO Fletcher
CO Fletcher
CO Fletcher
CO Fletcher
CO Lellouch
CO Lellouch
CO Lellouch
CO Lellouch
Residuals for CO Fletcher
Residuals for CO Fletcher
Residuals for CO Fletcher
Residuals for CO Fletcher
Residuals for CO Fletcher
Residuals for CO Lellouch
Residuals for CO Lellouch
Residuals for CO Lellouch
Residuals for CO Lellouch
Residuals for CO Lellouch
155
155
155
155
155
155
160
160
160
160
160
160
165
165
165
165
165
165
175
175
175
175
175
175
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
Wavelength (micrometer)
170
170
170
170
170
170
180
180
180
180
180
180
185
185
185
185
185
185
Fig. 6. CO lines at 153-187 µm, compared with models using
the CO distributions of Lellouch et al. (2005) and Fletcher et al.
(2010), shown in Fig. 2. CO lines occur at 154, 163, 174 and 186
µm. Other features are due to CH4 and H2O. The bottom curves
are difference (observed – modeled) plots (shifted by 0.97), fa-
voring the Lellouch et al. profile.
)
)
)
)
1
1
1
1
-
-
-
-
m
m
m
m
c
c
c
c
x
x
x
x
m
m
m
m
u
u
u
u
u
u
u
u
n
n
n
n
i
i
i
i
t
t
t
t
n
n
n
n
o
o
o
o
c
c
c
c
%
%
%
%
(
(
(
(
a
a
a
a
e
e
e
e
r
r
r
r
a
a
a
a
e
e
e
e
n
n
n
n
l
l
l
l
i
i
i
i
l
l
l
l
d
d
d
d
e
e
e
e
e
e
e
e
d
d
d
d
o
o
o
o
M
M
M
M
10
10
10
10
8
8
8
8
6
6
6
6
4
4
4
4
2
2
2
2
0
0
0
0
0
0
0
0
Profile A (disp. = 0.366)
Profile B (disp. = 0.464)
Uniform (disp. = 0.458)
8
2
2
2
2
8
8
8
Observed line area (% continuum x cm-1)
Observed line area (% continuum x cm-1)
Observed line area (% continuum x cm-1)
Observed line area (% continuum x cm-1)
4
4
4
4
6
6
6
6
Acknowledgements. PACS has been developed by a consortium of insti-
tutes led by MPE (Germany) and including UVIE (Austria); KUL, CSL,
IMEC (Belgium); CEA, OAMP (France); MPIA (Germany); IFSI, OAP/AOT,
OAA/CAISMI, LENS, SISSA (Italy); IAC (Spain). This development has been
supported by the funding agencies BMVIT (Austria), ESA-PRODEX (Belgium),
CEA/CNES (France), DLR (Germany), ASI (Italy), and CICT/MCT (Spain).
Additional funding support for some instrument activities has been provided by
ESA. Data presented in this paper were analysed using "HIPE", a joint develop-
ment by the Herschel Science Ground Segment Consortium, consisting of ESA,
the NASA Herschel Science Center, and the HIFI, PACS and SPIRE consortia.
We are indebted to Bruno B´ezard for important discussions on the HD and CH4
line parameters.
10
10
10
10
Fig. 5. Modeled vs. observed H2O line integrated areas for
the three water profiles of Fig. 2. Line areas are expressed in
cm−1 × % of the local continuum. For each profile, the mean
rms dispersion (in the same unit) between observed and mod-
eled areas is given. Profile A provides a better fit to the data than
do the other two profiles.
Recent CO observations at millimeter/submillimeter wave-
lengths (Lellouch et al. 2005, Hesman et al. 2007) point to a
higher abundance of CO in Neptune's stratosphere than in the
troposphere. Both studies thus indicate a dual external/internal
source, with the external source possibly provided by an ancient
cometary impact. They also provide consistent values of the CO
tropospheric mixing ratio (0.5-0.6 ppm). However, they differ by
References
Baines, K. & Hammel, H. 1994, Icarus, 109, 20
B´ezard, B., Romani, P., Feuchtgruber, H. & Encrenaz, T, 1998, Ap. J., 515, 868
B´ezard, B., Encrenaz, T., Lellouch, E., Feuchtgruber, H., 1999a, Science, 283,
800
B´ezard, B., Encrenaz, T, & Feuchtgruber, H., 1999b, ESA-SP 427, 153
Boudon, V., et al., 2010, J. Quant. Spectro. Rad. Transf., 111, 1117
Burgdorf, M., et al., 2003, Icarus, 164, 244
Conrath, B.J., Gierasch, P.J, & Ustinov, E.A., 1998, Icarus, 135, 501
Feuchtgruber, H., et al., 1997, Nature, 389, 159
Feuchtgruber, H., et al., 1999, A&A, 341, L17
Hammel, H.B., et al., ApJ, 644, 1326
Hartogh, P., et al. 2009, Planet. and Space Sci., 57, 1596
Hesman, B.E., Davis, G.R., Matthews, H.E., & Orton, G.S., 2007, Icarus, 186,
342
Lellouch, E., et al., 2001, A& A, 370, 610
Lellouch et al.: PACS observations of Neptune
5
Lellouch, E., Moreno, & G. Paubert, 2005, A& A, 430, L37
Lindal, G.F., 1992, AJ, 103, 967
Fletcher, L., et al. 2010, A&A, in press
Marten, A., et al., 2005, A&A, 1097
Meadows, V.S., et al., 1999, Icarus, 197, 585
Moses, J.I. et al., 2005, JGR, 110, E08001, doi:10.1029/2005JE002411
Orton, G.S., et al., 2007, A&A, 473, L5
Pilbratt, G., & many others 2010, A&A this issue
Poglitsch, A., Waelkens, C., Geis, N., Feuchtgruber, H., et al., 2010, A&A this
issue
PACS
Observers
Manual,
2010,
http:// Herschel.esac.esa.int/
Docs/PACS/pdf/pacs om.pdf
1 LESIA, Observatoire de Paris, 5 place Jules Janssen, F-92195
Meudon, France
e-mail: [email protected]
2 Max-Planck-Institut
fur Sonnensystemforschung, Katlenburg-
Lindau, Germany
3 Max-Planck-Institut
Giessenbachstrasse, 85748 Garching, Germany
4 Instituut voor Sterrenkunde, Katholieke Universiteit Leuven,
fur
extraterrestrische
Physik,
5 SRON, Groningen, the Netherlands
6 Jet Propulsion Laboratory, California Institute of Technology,
Pasadena, United States
7 Space Research Centre, Polish Academy of Science, Warszawa,
Belgium
Poland
8 Blue Sky Spectroscopy Inc., Lethbridge, Alberta, Canada
9 European Space Astronomy Center, Madrid, Spain
10 Instituto de Astrof´ısica de Andaluc´ıa (CSIC), Granada, Spain
11 California Institute of Technology, Pasadena, United States
12 Rutherford Appleton Laboratory, Oxfordshire, United Kingdom
13 Deutsches Zentrum fur Luft- und Raumfahrt
(DLR), Bonn,
Germany
14 University of Michigan, Ann Arbor, United States
15 Universit´e de Bordeaux, Observatoire Aquitain des Sciences de
l'Univers, CNRS, UMR 5804, Laboratoire d'Astrophysique de
Bordeaux, France
16 Laboratorio de Astrof´ısica Molecular, CAB. INTA-CSIC, Spain
17 Joint Astronomy Center, Hilo, United States
18 LERMA, Observatoire de Paris, and Univ. Pierre et Marie Curie,
19 F.R.S.-FNRS, Institut d'Astrophysique et de G´eophysique, Li`ege,
20 Herschel Science Centre, European Space Astronomy Centre,
21 University of Lethbridge, Canada
22 University of Bern, Switzerland
23 University of Cologne, Germany
24 Max-Planck-Institut fr Radioastronomie, Bonn, Germany
25 Experimental Physics Dept., National University of
Maynooth, Co. Kildare. Ireland
Ireland
Paris, France
Belgium
Madrid, Spain
|
1011.2229 | 1 | 1011 | 2010-11-09T22:31:04 | System parameters, transit times and secondary eclipse constraints of the exoplanet systems HAT-P-4, TrES-2, TrES-3 and WASP-3 from the NASA EPOXI Mission of Opportunity | [
"astro-ph.EP"
] | As part of the NASA EPOXI Mission of Opportunity, we observed seven known transiting extrasolar planet systems in order to construct time series photometry of extremely high phase coverage and precision. Here we present the results for four "hot-Jupiter systems" with near-solar stars - HAT-P-4, TrES-3, TrES-2 and WASP-3. We observe ten transits of HAT-P-4, estimating the planet radius Rp = 1.332 \pm 0.052 RJup, the stellar radius R \star = 1.602 \pm 0.061 R \odot, the inclination i = 89.67 \pm 0.30 degrees and the transit duration from first to fourth contact T = 255.6 \pm 1.9 minutes. For TrES-3, we observe seven transits, and find Rp = 1.320 \pm 0.057 RJup, R\star = 0.817 \pm 0.022 R\odot, i = 81.99 \pm 0.30 degrees and T = 81.9 \pm 1.1 minutes. We also note a long term variability in the TrES-3 light curve, which may be due to star spots. We observe nine transits of TrES-2, and find Rp = 1.169 \pm 0.034 RJup, R\star = 0.940 \pm 0.026 R\odot, i = 84.15 \pm 0.16 degrees and T = 107.3 \pm 1.1 minutes. Finally we observe eight transits of WASP-3, finding Rp = 1.385 \pm 0.060 RJup, R\star = 1.354 \pm 0.056 R\odot, i = 84.22 \pm 0.81 degrees and T = 167.3 \pm 1.3 minutes. We present refined orbital periods and times of transit for each target. We state 95% confidence upper limits on the secondary eclipse depths in our broadband visible bandpass centered on 650 nm. These limits are 0.073% for HAT-P-4, 0.062% for TrES-3, 0.16% for TrES-2 and 0.11% for WASP-3. We combine the TrES-3 secondary eclipse information with the existing published data and confirm that the atmosphere likely does not have a temperature inversion. | astro-ph.EP | astro-ph | System parameters, transit times and secondary eclipse constraints of the
exoplanet systems HAT-P-4, TrES-2, TrES-3 and WASP-3 from the NASA
EPOXI Mission of Opportunity.
Jessie L. Christiansen1, Sarah Ballard1, David Charbonneau1, Drake Deming2,
Matthew J. Holman1, Nikku Madhusudhan3, Sara Seager3, Dennis D. Wellnitz4,
Richard K. Barry2, Timothy A. Livengood2, Tilak Hewagama2,4, Don L. Hampton5,
Carey M. Lisse6, and Michael F. A'Hearn4
ABSTRACT
As part of the NASA EPOXI Mission of Opportunity, we observed seven known
transiting extrasolar planet systems in order to construct time series photometry of
extremely high phase coverage and precision. Here we present the results for four "hot-
Jupiter systems" with near-solar stars -- HAT-P-4, TrES-3, TrES-2 and WASP-3. We
observe ten transits of HAT-P-4, estimating the planet radius Rp = 1.332± 0.052 RJup,
the stellar radius R⋆ = 1.602 ± 0.061 R⊙, the inclination i = 89.67 ± 0.30 degrees and
the transit duration from first to fourth contact τ = 255.6 ± 1.9 minutes. For TrES-3,
we observe seven transits, and find Rp = 1.320 ± 0.057 RJup, R⋆ = 0.817 ± 0.022 R⊙,
i = 81.99±0.30 degrees and τ = 81.9±1.1 minutes. We also note a long term variability
in the TrES-3 light curve, which may be due to star spots. We observe nine transits of
TrES-2, and find Rp = 1.169 ± 0.034 RJup, R⋆ = 0.940 ± 0.026 R⊙, i = 84.15 ± 0.16
degrees and τ = 107.3 ± 1.1 minutes. Finally we observe eight transits of WASP-3,
finding Rp = 1.385 ± 0.060 RJup, R⋆ = 1.354 ± 0.056 R⊙, i = 84.22 ± 0.81 degrees and
τ = 167.3± 1.3 minutes. We present refined orbital periods and times of transit for each
target. We state 95% confidence upper limits on the secondary eclipse depths in our
broadband visible bandpass centered on 650 nm. These limits are 0.073% for HAT-P-4,
0.062% for TrES-3, 0.16% for TrES-2 and 0.11% for WASP-3. We combine the TrES-3
secondary eclipse information with the existing published data and confirm that the
atmosphere likely does not have a temperature inversion.
1Harvard-Smithsonian Center
for Astrophysics,
60 Garden Street, Cambridge, MA 02138, USA;
[email protected]
2Goddard Space Flight Center, Greenbelt, MD 20771, USA
3Massachusetts Institute of Technology, Cambridge, MA 02159, USA
4University of Maryland, College Park, MD 20742, USA
5University of Alaska Fairbanks, Fairbanks AK 99775, USA
6Johns Hopkins University Applied Physics Laboratory, Laurel, MD 20723, USA
0
1
0
2
v
o
N
9
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
2
2
2
.
1
1
0
1
:
v
i
X
r
a
-- 2 --
Subject headings: planetary systems -- eclipses -- stars: individual (HAT-P-4, WASP-
3, TrES-2, TrES-3)
1.
Introduction
The EPOXI Mission of Opportunity is a re-purposing of the Deep Impact flyby spacecraft, and
comprises the Extrasolar Planet Observation and Characterization (EPOCh) investigation and the
Deep Impact eXtended Investigation (DIXI). The primary goal of EPOCh was to scrutinize a
small set of known transiting extrasolar planets. From 2008 January to 2008 August, we used the
high resolution imaging (HRI) instrument (Hampton et al. 2005) and a broadband visible filter
to construct high precision, high phase coverage and high cadence light curves for seven targets.
We observed each target nearly continuously for several weeks at a time. The main science goals
of EPOCh were to refine the system parameters of the known planets, to search for additional
planets both directly (via transits of the additional body) and indirectly (via induced changes in
the transits of the known planet), and to constrain the reflected light from the known planet at
secondary eclipse. It is also useful to provide updated periods and times of epoch for these systems
in order to reduce uncertainties on predicted transit and eclipse times, and therefore maximize
the return of follow-up observations. In previous EPOCh papers we have presented the search for
additional planets in the GJ 436 system (Ballard et al. 2010) and the secondary eclipse constraints
for HAT-P-7 (Christiansen et al. 2010). In this paper we present the updated system parameters,
including constraints on the transit timing and changes in the transit parameters, and secondary
eclipse constraints for a further four targets: HAT-P-4, TrES-3, TrES-2 and WASP-3, introduced
below. The search for additional planets in these systems will be presented in a separate paper
(Ballard et al. in prep).
The exoplanet HAT-P-4b (Kovacs et al. 2007), orbits a slightly evolved metal-rich late F star.
With a mass of 0.68 MJup and a radius of 1.27 RJup, it joined the ranks of inflated planets that
have continued to challenge models of the physical structure of hot Jupiters.
TrES-3 (O'Donovan et al. 2007) is notable for its very short orbital period of 1.30619 days.
This proximity to the star makes TrES-3 a promising target for observations of reflected light at
visible wavelengths; the planet-to-star flux ratio as measured in reflected light during the secondary
eclipse is given by Ag(Rp/a)2, where Ag is the geometric albedo, Rp is the planetary radius and
a is the semi-major axis of the planetary orbit. Winn et al. (2008), de Mooij & Snellen (2009)
and Fressin et al. (2010) have observed secondary eclipses of TrES-3 at visible and near-infrared
wavelengths, and the emerging picture of the planetary atmosphere is one with efficient day-night
re-circularization and no temperature inversion in the upper atmosphere. This is in contrast to
predictions of a temperature inversion based on the high level of irradiation (Fortney et al. 2008).
Sozzetti et al. (2009) studied the transit timing variations of TrES-3 and noted significant outliers
-- 3 --
from a constant period. Gibson et al. (2009) monitored further transit times of TrES-3 and ruled
out sub-Earth mass planets in the exterior and interior 2:1 resonances for circular orbits.
TrES-2 (O'Donovan et al. 2006) was the first transiting planet found in the field of view of
the NASA Kepler mission (Borucki et al. 2009). Holman et al. (2007) noted that the high impact
parameter (b ≈ 0.85) of TrES-2 made transit parameters such as inclination and duration sensitive
to changes due to orbital precession. Mislis & Schmitt (2009) and Mislis et al. (2010) claimed a
significantly shorter duration for TrES-2 transits two years after the measurements of Holman et al.
(2007). They proposed that this was caused by a change in orbit inclination due to precession, and
that the duration would continue to decrease. However, Scuderi et al. (2009) measured a duration
consistent with O'Donovan et al. (2006) and Holman et al. (2007) and did not see the predicted
trend of decreasing transit duration. Secondary eclipses of TrES-2 have been observed in the near-
infrared (O'Donovan et al. 2010), and the results favor a thermal inversion in the upper atmosphere,
supporting the hypothesis that highly irradiated planetary atmospheres have inversions. The transit
timing variations of TrES-2 have been studied by Raetz et al. (2009) and Rabus et al. (2009), who
find no statistically significant variations.
WASP-3b (Pollacco et al. 2008), with a short period (1.84634 days) and a hot host star (F7-
F8V, Teff = 6400K), is one of the hottest transiting planets known, and another very good target
for observing reflected light at secondary eclipse.
The paper is organized as follows. The observations and generation of the light curves are
described in Section 2, the transit analysis is presented in Section 3, the secondary eclipse analysis
is presented in Section 4 and the results are discussed in Section 5.
2. Observations and analysis
The EPOCh observations were made using the high resolution imager (HRI), which has a 30-cm
aperture and a 1024×1024 pixel CCD. For our observations we used a clear visible filter, covering
350 -- 1000nm, in order to maximize the throughput of photons. The integration time for the science
observations was 50 seconds, which for near-continuous observations results in roughly 1500 images
per day. Since the on-board spacecraft memory is only 300Mb, we initially chose to read out only
a 128 × 128 pixel sub-array of the full CCD, to ensure full phase coverage between data downlinks
from the spacecraft. The CCD comprises four quadrants that are read out independently, and the
sub-array is centered on the CCD where the four quadrants meet. The pixel scale is 0.4 arcsec
pixel−1, resulting in a sub-array field of view of 0.72 square arcminutes. The images are significant
defocused, resulting in a stellar point-spread function (PSF) with a full-width half maximum of 4
arcseconds. Typically this meant that the target star was the only star in the field of view, and
we were unable to employ relative photometry techniques for removing correlated noise in the light
curves.
Table 1 summarizes the observing schedules for each of the four targets. HAT-P-4 and TrES-3,
-- 4 --
along with GJ436 and XO-2, were observed during the initial observing block from 2008 January
to 2008 May. The project was awarded an additional contingent observing block from 2008 June
to 2008 August, during which time HAT-P-4 was re-observed, and TrES-2 and WASP-3 were also
observed. During the contingent observations we began observing in a larger 256 × 256 pixel sub-
array mode, to reduce losses from pointing drifts that occasionally resulted in the target star lying
outside of the 128× 128 pixel sub-array field of view. The number of images that could be obtained
with the larger sub-array mode between data downlinks from the spacecraft was constrained by
the data storage capacity on board the spacecraft. Therefore, in order to maximize the phase
coverage we chose to restrict observations in the 256 × 256 pixel sub-array mode to the times of
particular interest -- during the transits and secondary eclipses. One event per data downlink could
be observed in the larger sub-array mode without reducing the temporal coverage. Table 1 shows
the total number of transits and eclipses observed for each target, with the number observed in the
256× 256 pixel sub-array mode given in parentheses. As discussed in Section 1, TrES-2 was claimed
to show changes in the transit inclination with time. Therefore, we used the larger sub-array mode
to observe the transits of TrES-2 where possible. WASP-3 was a promising target for secondary
eclipse observations, and therefore we observed the secondary eclipses of WASP-3 in the larger
mode where possible. For HAT-P-4 we observed two of the three transits and two of the three
eclipses obtained in the contingent observations in the 256× 256 pixel sub-array mode. TrES-3 was
observed in the initial observing block and no observations were obtained in the larger mode.
2.1.
Image Calibration and Time Series Extraction
We receive calibrated FITS images from the extant Deep Impact data reduction pipeline
(Klaasen et al. 2005). These data have been bias- and dark-subtracted and flat-fielded, using
calibration images obtained on the ground before launch. Due to the very high precision required
in the light curves, we perform several additional calibration steps to account for changes in the
CCD since launch. The spacecraft pointing drifts considerably with time, resulting in significant
coverage of the CCD by the stellar PSF and placing paramount importance on the flat-fielding.
The procedure is described in Ballard et al. (2010) and summarized here.
Table 1. EPOCh observations
Target
V Mag
UT Dates observed (2008)
No. of Transitsa No. of Eclipsesa
HAT-P-4
TrES-3
TrES-2
WASP-3
11.22
11.18
11.41
10.64
01/22 -- 02/12, 06/29 -- 07/07
03/06 -- 03/18
06/27 -- 06/28, 07/19 -- 07/29
07/17 -- 07/18, 07/30 -- 08/07, 08/10 -- 08/15
10 (2)
7 (0)
9 (7)
8 (0)
9 (2)
6 (0)
8 (2)
9 (8)
aIncluding partial events. The number in brackets is the subset of events observed in 256 × 256 pixel
sub-array mode.
-- 5 --
For each target, we use a PSF constructed from the images to locate the position of the star to
a hundredth of a pixel. At this stage we reject images with 10σ outliers from the PSF fit, assuming
the stellar PSF to be contaminated by an energetic particle hit. We subtract a time-dependent bias
calculated for each quadrant from the corresponding overscan region. We reduce the pixels in the
central columns and rows of the CCD (forming the internal boundaries between the quadrants) by
roughly 15% and 1% respectively, to correct an artifact produced by the CCD readout electronics.
For data obtained in the 256×256 pixel mode, we scale the images by a constant (typically differing
from unity by one part in a thousand) to correct an observed flux offset between the two sub-array
modes.
In order to track time-dependent changes in the flat-field since launch, there is a small green
LED stimulation lamp that can be switched on to illuminate the CCD. We obtained blocks of 200
calibration frames using this lamp, which were taken every few days throughout the observations,
alternating between blocks in the smaller and larger sub-array modes in the contingent observations.
We correct each science frame by the flat-field generated from lamp images taken in the same sub-
array mode. We assume any remaining flat-field errors to be color-dependent and therefore unable
to be addressed by the monochromatic lamp.
We perform aperture photometry, using a circular aperture of radius 10 pixels. The resulting
light curves exhibit significant correlated noise on the order of 1%, which is associated with the
drift in the spacecraft pointing. In order to correct for this, we use the data itself to generate a
sensitivity map of the CCD. We assume the out-of-transit and out-of-eclipse data to be of uniform
brightness, with two caveats. First, the star may have intrinsic variations in stellar brightness due
to spots. Only one of the four targets displayed long-period variability (Figure 3), and this was
removed by fitting and removing a polynomial in time before producing the CCD sensitivity map.
Second, transits of additional planets may be present, which will be suppressed with this treatment
(Ballard et al. 2010). We randomly draw several thousand of the out-of-transit and out-of-eclipse
points and find a robust average flux of the 30 spatially nearest neighbors. We use this set of
averages to generate a two-dimensional surface spline to the flux distribution across the CCD. Each
point in the light curve is then corrected by interpolating onto this surface. The entire procedure
is iterated several times to converge on the positions and scaling factors that result in the lowest
scatter in the out-of-transit and out-of-eclipse data in the final light curve.
The robustness of the surface spline for each target depends on the coverage of the CCD by
that target. If the coverage is small and the corresponding density of photometry apertures high,
then there is a high probability that the same pixel will be returned to multiple times over the
observations. Having flux measurements separated in time reduces the influence of stellar activity
on our calibration of the sensitivity of each pixel. Figure 1 shows the complete CCD coverage for
two targets. TrES-2 is well confined on the CCD and the density of photometry apertures leads
to a more robust surface spline. The TrES-2 light curve prior to and post the application of the
surface spline is shown in Figure 4. On the other hand, the photometry apertures for WASP-3
sample a much larger area of the CCD, and in addition many of the observations obtained in the
-- 6 --
256×256 pixel sub-array mode do not overlay the central 128×128 pixel sub-array. The resulting
surface spline is therefore more sensitive to noise introduced by stellar activity or systematics that
are not an artifact of the pointing jitter. The WASP-3 light curve prior to and post the application
of the surface spline is shown in Figure 5. The lower panel of Figure 5 shows how the noise in the
final calibrated WASP-3 light curve bins down compared with the expectation for Gaussian noise,
and the poor quality of the data is due to the low density of the CCD coverage for WASP-3.
Fig. 1. -- The CCD positions of the photometry apertures for two targets. Left: TrES-2 is confined
to the center of the CCD and therefore the same pixels are sampled well for creating a robust surface
spline. Right: WASP-3 is spread over a much larger fraction of the CCD, including large excursions
out of the central 128×128 pixel sub-array when observations were obtained in the 256×256 pixel
sub-array. This reduces the quality of the surface spline and results in a larger component of
correlated noise in the WASP-3 light curve.
2.2. Details for each target
The final HAT-P-4 light curve is shown in Figure 2. HAT-P-4 was the first EPOCh target
observed, initially for 22 days from 2008 January 22 to 2008 February 12, during the original
EPOCh target schedule, and again for 8 days from 2008 June 29 to 2008 July 7 during the contingent
observations. Of the 45,320 images obtained of HAT-P-4, 5434 were discarded due to the star being
either out of the field of view or too close to the edge of the CCD to measure accurate photometry,
1305 were discarded due to energetic particle hits, and 76 were discarded due to readout smear,
for a final total of 38,505 acceptable images. All of the data obtained in the initial run are in the
128 × 128 pixel sub-array mode. Of the contingent data, two of the three transits and two of the
-- 7 --
three eclipses are in the larger 256 × 256 pixel sub-array mode, and the remaining data are in the
smaller mode. The bottom panel of Figure 2 shows how the scatter in the final light curve scales
down with increasing bin size -- for Gaussian noise the expectation is the scatter will decrease as
1/√N , where N is the number of points in the bin.
Fig. 2. -- Upper panel: The full HAT-P-4 EPOXI light curve. The left panel shows the original
run of seven consecutive transits. The right panel shows the three transits observed five months
later during the EPOCh contingent observations. In each panel the lower curve is before the first
application of the surface spline and the upper curve is the final calibrated light curve. The red
data points were obtained in the larger 256 × 256 pixel sub-array mode. Lower panel: The scatter
in the out-of-transit data with increasing bin size (diamonds) and comparing to the expectation
for Gaussian noise (1/√N , where N is the number of points in the bin, shown as the solid line
normalized to the unbinned value of the scatter). The points do not follow the line, indicating
residual correlated noise in the light curve.
The final TrES-3 light curve is shown in Figure 3. TrES-3 was the second EPOCh target
observed, for 12 days from 2008 March 6 to 2008 March 18. The gap in the light curve from 2 -- 5
days is due to a 'pre-look' for the subsequent EPOCh target, XO-2, which was performed in order
to refine the pointing for that target. We obtained a total of 14,195 images of TrES-3, of which
we discarded 1165 due to the star being out of or too close to the edge of the field of view, 1632
-- 8 --
due to energetic particle hits and 127 due to readout smear, leaving 11,271 images. We obtained
all of the TrES-3 data in the 128 × 128 pixel sub-array mode. After the initial application of the
two-dimensional surface spline a long timescale, low amplitude variability was evident in the light
curve. This can be seen in the lower light curve in Figure 3. In order to remove this variability we
bin the out-of-transit data by two hours and fit with a time-dependent fifth-order polynomial for
the data occurring later than 4.0 days. We divide out this feature before iterating over the previous
steps to produce the final light curve. The polynomial is plotted on the lower light curve, and the
final light curve is shown as the upper curve in Figure 3. As with HAT-P-4, the bottom panel of
Figure 3 shows the noise properties of the data.
Fig. 3. -- Upper panel: The TrES-3 EPOXI light curve. The gap from 2.5 -- 5 days is during the
pre-look for a subsequent target. Seven transits of TrES-3 were observed in total. The lowest
light curve is prior to the first application of the surface spline, the middle light curve is after the
application of the spline but prior to the removal of the time-dependent polynomial, and the upper
light curve is the final calibrated data set. Lower panel: See Figure 2 for explanation. In the case
of TrES-3, where all data were obtained in the smaller sub-array mode and the total time span is
relatively short, the scatter bins down close to the expectation for Gaussian noise.
The final TrES-2 light curve is shown in Figure 4. We observed TrES-2 during the contingent
EPOCh observations, from 2008 July 7 to 2008 July 30, in addition to a pre-look for pointing on
-- 9 --
2008 June 28 and 29. In total, we obtained 31,210 images of TrES-2, with 1979 discarded due to
the star lying out of or too close to the edge of the field of view, 1427 discarded due to energetic
particle hits and 80 discarded due to readout smear, for a total of 27,724 acceptable images. We
observed nine transits in total, including seven in the 256 × 256 pixel sub-array mode. The lower
panel of Figure 4 shows that correlated noise remains in the final light curve.
Fig. 4. -- Upper panel: The TrES-2 EPOXI light curve. The data obtained from days 1 -- 3 are the
pre-look, for refinement of the spacecraft pointing. From days 3 -- 11 the spacecraft was observing a
different target before returning to TrES-2 with updated pointing parameters. The gap from days
21 -- 23 spans the pre-look for the subsequent target. Nine transits of TrES-2 were observed in total.
The lower curve is prior to the first application of the surface spline and the upper curve is the final
calibrated light curve. The red data points were obtained in the larger 256 × 256 pixel sub-array
mode. Lower panel: See Figure 2 for explanation.
The final EPOCh light curve for WASP-3 is shown in Figure 5. We observed WASP-3 during
the contingent observations, from 2008 July 29 to 2008 August 16, with a pre-look from 2008 July
17 to 2008 July 19. We obtained 24,015 images of WASP-3, of which we discarded 4,182 due to
the star being out of or too close to the edge of the field of view, 403 due to energetic particle
hits, and 808 due to readout smear, leaving 18,622 acceptable images. For WASP-3, none of the
eight transits were observed in 256× 256 pixel sub-array mode, however eight of the nine secondary
-- 10 --
eclipses were observed in this mode. The two-dimensional surface spline relies on multiple visits to
the same part of the CCD to characterize robustly the interpixel variations. This is particularly
true for the data that occur during the transits and eclipses, since they cannot be assumed to be
of uniform flux and are therefore excluded from the creation of the surface. In order to effectively
flat-field the data that are taken during transit and eclipse, the observations taken during these
times must be gathered at the same spatial positions as data obtained at other times. In the case of
WASP-3, four of the eight secondary eclipses occurred at locations that were poorly sampled. No
out-of-transit or out-of-eclipse observations fell on these pixels, and therefore we cannot estimate
the true sensitivity of these pixels in order to produce an effective flat-field. These eclipses occur
at 1.0, 17.6, 19.3 and 26.6 days, and can be seen in the light curve as increases in flux. These
four eclipses are discarded for the final analysis. Besides these events, a significant fraction of the
WASP-3 data are distributed in poorly-sampled areas of the CCD, degrading the robustness of the
two-dimensional surface spline. The bottom panel of Figure 5 demonstrates the adverse effect this
has on the noise properties of the final light curve, as the data do not bin down as expected for
Gaussian noise.
3. Transit analysis
For the transit analysis, we make several additional calibration steps. The two-dimensional
surface spline uses only a fraction of the data to generate the surface, in order to preserve as much
of the information in the light curve as possible, and to minimize the suppression of transits of
putative additional planets. However for the transit analysis, we use all of the available data to
calibrate each event. For each transit, we define a window approximately three times the duration
of the transit, centered on the predicted transit time. We take each point in this window and divide
the flux by a robust average of the 30 spatially nearest points that do not fall in any of the transit
windows. This is essentially a point-by-point application of the full two-dimensional surface spline.
We then fit a slope, linear with time, to the out-of-transit data across each transit and divide it
out, to remove any residual long timescale trends.
For TrES-3, TrES-2 and WASP-3, we generate non-linear limb-darkening coefficients of the
form given by Claret (2000), Iµ/I1 = 1 − P4
n=1 cn(1 − µn/2), where I1 is the specific intensity at
the center of the disk and µ = cos(γ), with γ the angle between the emergent intensity and the line
of sight. We use photon-weighted stellar atmosphere models of Kurucz (1994, 2005) that bracket
the published values of stellar Teff and log g, and convolve these with the total EPOXI response
function, including filter, optics and CCD response. We fit for the four coefficients of the non-linear
form of the limb-darkening using 17 positions across the stellar limb, at 2 nm intervals along the
350 -- 1000 nm bandpass. We calculate the final set of coefficients as the weighted average when
integrated over the bandpass, and bi-linearly interpolate across Teff and log g for each target. The
final set of coefficient for each targets is given in Table 3 for TrES-3, Table 4 for TrES-2 and Table
5 for WASP-3.
-- 11 --
Fig. 5. -- Upper panel: The WASP-3 EPOXI light curve. The first two days of data are the pre-look
to refine the pointing. The gap between 22 and 24 days is due to the pre-look for the subsequent
target. The significant positive deviations seen at 1.0, 17.6, 19.3 and 26.6 days are instrumental in
nature; see the text for details. The lower curve is prior to the first application of the spatial spline
and the upper curve is the final calibrated light curve. The red data points were obtained in the
larger 256 × 256 pixel sub-array mode. Lower panel: See Figure 2 for explanation.
The quality of the EPOCh light curves is nearly sufficient to fit for the limb-darkening coeffi-
cients rather than assuming theoretical values. Ultimately, the degeneracies between the geometric
parameters of the transiting system and the limb-darkening coefficients prevent us from placing
meaningful constraints on the coefficients. In the case of HAT-P-4 however, the system is very close
to edge-on (i = 89.9+0.1
−2.2 degrees), which reduces the parameter space considerably. Therefore, for
HAT-P-4 we instead use a quadratic equation for the limb-darkening, Iµ/I1 = 1−a(1−µ)−b(1−µ)2,
and allow two linear combinations of the coefficients, c1 = 2a + b and c2 = a − 2b to be free param-
eters in the transit analysis, which produced a better fit to the data as defined below.
When fitting the transits, we use the analytic equations of Mandel & Agol (2002) to generate
a model transit, and use χ2 as a goodness-of-fit estimator. We use the Levenberg-Marquardt
algorithm to fit three dimensionless geometric parameters of the system: Rp/R⋆, R⋆/a and cos i,
where Rp is the planetary radius, R⋆ is the stellar radius, a is the semi-major axis of the planetary
-- 12 --
orbit and i is the inclination of the orbit. We fix the period to the published value, but allow the time
of center of transit to vary independently for each of the transits. We then use the published mass
values for each of the systems to convert the transit parameters to physical properties, drawing
values from Kovacs et al. (2007) for HAT-P-4, Sozzetti et al. (2009) for TrES-3, Sozzetti et al.
(2007) for TrES-2 and Pollacco et al. (2008) for WASP-3. The final results of these fits are given
in Table 2 for HAT-P-4, Table 3 for TrES-3, Table 4 for TrES-2 and Table 5 for WASP-3. We
also give the transit duration from first to fourth contact for each best-fit model. For WASP-3, we
discard the final transit (which was significantly offset in flux due to correlated noise), and also a
partial transit (which included only the ingress), for a total of six transits. The phase-folded and
binned transits for each target are shown in Figure 6 for HAT-P-4, Figure 7 for TrES-3, Figure 8
for TrES-2 and Figure 9 for WASP-3.
The errors on the parameters are calculated using the residual permutation "rosary bead"
method (Winn et al. 2008). For each target, we find the residuals to the best-fit model. We shift
these residuals forward collectively to the next time stamp and add the best fit models back to the
new residuals, generating a new realization of the light curve which retains the correlated noise
signals in the original light curve. We repeat this process 8000 times (covering approximately six
days) and each time we fit for and record the geometric parameters, times of center of transit, and
limb-darkening coefficients where appropriate. For each parameter we construct a histogram of the
8000 measurements, to which we fit a Gaussian. We then define the error on that parameter by
the half-width half-maximum value of the best fit Gaussian. We find that increasing the number
of iterations beyond 4000 does not significantly change the calculated errors.
To find the errors in the transit times, we perform a second rosary bead analysis, holding
the geometric and limb-darkening values fixed and allowing only the times of center of transit to
vary. We find that 4000 iterations are sufficient to sample the range of correlated noise signals, and
calculate the errors in the same fashion as the geometric parameters. For each target we calculate
a new orbital period and epoch by performing a weighted linear fit to the EPOCh transit times
and any published transit times.
4. Secondary eclipse constraints
Our constraints of the secondary eclipse depths are limited by the correlated noise in the
EPOXI data. Ideally, for each target we would combine our multiple observations of the secondary
eclipses to increase the signal to noise. However, the fluctuations due to the correlated noise
preclude this. For example, Figure 11 shows six of the TrES-3 secondary eclipses, where in some
cases correlated noise results in an increase in flux at the time of secondary eclipse, instead of
the expected decrement. If we assume that the secondary eclipse in the EPOCh bandpass, with a
central wavelength of 650 nm, is due exclusively to the reflected light of the planet, then the eclipse
depths we would anticipate, for a geometric albedo of 1, would range from 0.02% for HAT-P-4 to
-- 13 --
Table 2. HAT-P-4 system parameters
Parameter
Value
Adopted valuesa
M⋆ (M⊙)
Mp (MJup)
Transit fit values
Rp/R⋆
a/R⋆
i (deg)
Derived values
P (days)
Tc (BJD)
R⋆ (R⊙)
Rp (RJup)
τ (mins)
Limb-darkening coefficients
a
b
Transit times (BJD)
1.26 ± 0.14
0.68 ± 0.04
0.0855 ± 0.0078
0.1672 ± 0.0078
89.67 ± 0.30
3.0565114 ± 0.0000028
2, 454, 502.56227 ± 0.00021
1.602 ± 0.061
1.332 ± 0.052
255.6 ± 1.9
0.314
0.366
2, 454, 490.33445 ± 0.00072
2, 454, 493.39232 ± 0.00061
2, 454, 496.44984 ± 0.00056
2, 454, 499.50426 ± 0.00070
2, 454, 502.56156 ± 0.00056
2, 454, 505.62006 ± 0.00082
2, 454, 508.67569 ± 0.00056
2, 454, 649.27624 ± 0.00064
2, 454, 652.33053 ± 0.00065
2, 454, 655.38842 ± 0.00065
aMasses are from Kovacs et al. (2007).
-- 14 --
Table 3. TrES-3 system parameters
Parameter
Value
Adopted valuesa
M⋆ (M⊙)
Mp (MJup)
Transit fit values
Rp/R⋆
a/R⋆
i (deg)
Derived values
P (days)
Tc (BJD)
R⋆ (R⊙)
Rp (RJup)
i (deg)
τ (mins)
Limb-darkening coefficients
c1
c2
c3
c4
Transit times (BJD)
0.928+0.028
−0.048
1.910+0.075
−0.080
0.1661 ± 0.0343
0.1664 ± 0.0204
81.99 ± 0.30
1.30618608 ± 0.00000038
2, 454, 538.58069 ± 0.00021
0.817 ± 0.022
1.320 ± 0.057
81.99 ± 0.30
81.9 ± 1.1
0.5169
-0.6008
1.4646
-0.5743
2, 454, 532.04939 ± 0.00033
2, 454, 533.35515 ± 0.00035
2, 454, 537.27463 ± 0.00038
2, 454, 538.58126 ± 0.00035
2, 454, 539.88703 ± 0.00040
2, 454, 541.19261 ± 0.00035
2, 454, 542.49930 ± 0.00041
aMasses are from Sozzetti et al. (2009).
-- 15 --
Table 4. TrES-2 system parameters
Parameter
Value
Adopted valuesa
M⋆ (M⊙)
Mp (MJup)
Transit fit values
Rp/R⋆
a/R⋆
i (deg)
Derived values
P (days)
Tc (BJD)
R⋆ (R⊙)
Rp (RJup)
τ (mins)
Limb-darkening coefficients
c1
c2
c3
c4
Transit times (BJD)
0.98 ± 0.062
1.198 ± 0.053
0.1278 ± 0.0094
0.1230 ± 0.0179
84.15 ± 0.16
2.47061344 ± 0.0000075
2, 454, 4664.23039 ± 0.00018
0.940 ± 0.026
1.169 ± 0.034
107.3 ± 1.1
0.3899
-0.1391
0.9662
-0.0329
2, 454, 646.93735 ± 0.00032
2, 454, 656.81879 ± 0.00034
2, 454, 659.28871 ± 0.00042
2, 454, 661.76005 ± 0.00044
2, 454, 664.23072 ± 0.00050
2, 454, 669.17156 ± 0.00028
2, 454, 671.64117 ± 0.00028
2, 454, 674.11318 ± 0.00033
2, 454, 676.58257 ± 0.00051
aMasses are from Sozzetti et al. (2007).
-- 16 --
Table 5. WASP-3 system parameters
Parameter
Value
Adopted valuesa
M⋆ (M⊙)
Mp (MJup)
Transit fit values
Rp/R⋆
a/R⋆
i (deg)
Derived values
P (days)
Tc (BJD)
R⋆ (R⊙)
Rp (RJup)
i (deg)
τ (mins)
Limb-darkening coefficients
c1
c2
c3
c4
Transit times (BJD)
1.24+0.06
−0.11
1.76+0.08
−0.14
0.1051 ± 0.0124
0.1989 ± 0.0287
84.15 ± 0.16
1.8468373 ± 0.0000014
2, 454, 686.82069 ± 0.00039
1.354 ± 0.056
1.385 ± 0.060
84.22 ± 0.81
167.3 ± 1.3
0.2185
0.6183
-0.1040
-0.0426
2, 454, 679.43264 ± 0.00050
2, 454, 681.27911 ± 0.00040
2, 454, 683.12740 ± 0.00035
2, 454, 684.97486 ± 0.00027
2, 454, 686.82053 ± 0.00059
2, 454, 690.51381 ± 0.00055
2, 454, 692.36117 ± 0.00043
2, 454, 694.20711 ± 0.00042
aMasses are from Pollacco et al. (2008).
-- 17 --
Fig. 6. -- The seven HAT-P-4 transits from the original observing schedule, phase-folded and binned
in five minute intervals. light curve. The solid line is the best fit transit model. Lower panel: The
residuals when the best-fit model is subtracted from the data.
0.08% for TrES-3.1
Since the fluctuations from correlated noise in the measured eclipse depths are sometimes
larger than the signal we expect to measure, we choose not to combine the multiple observations
and instead analyze each eclipse independently. Our intent is to use the scatter of individual eclipse
measurements to constrain the amplitude of the correlated noise. As for the transits, for each eclipse
in the data we apply a point-by-point correction to the data in and adjacent to the eclipse. For the
targets presented here we assume that e = 0 and therefore that the secondary eclipse occurs at a
phase of 0.5. For TrES-3 and TrES-2 this assumption is strongly supported by previous secondary
eclipse measurements with the Spitzer IRAC instrument which demonstrated no evidence of non-
zero eccentricity (Fressin et al. 2010 and O'Donovan et al. 2010 respectively). For HAT-P-4 and
WASP-3, the extant radial velocity data are consistent with circular orbits (Kovacs et al. 2007
1In fact, the CCD is found to be quite efficient at the redder wavelengths, and it is therefore feasible that for the
hottest planets there may be a contribution from the thermal emission of the planet, resulting in deeper secondary
eclipses.
-- 18 --
Fig. 7. -- Upper panel: The seven TrES-3 transits, phase-folded and binned in two minute intervals.
The solid line is the best-fit transit model. Lower panel: The residuals when the best-fit model is
subtracted from the data.
and Pollacco et al. 2008 respectively). In addition to the point-by-point correction, we fit a linear
time-dependent slope to the adjacent out-of-eclipse data to remove any remaining long timescale
trends. Finally, we separate the data into 10 minute bins and remove 3σ flux outliers from each
bin. For TrES-2, we discard the first observed eclipse, which was obtained during the pre-look for
this target, since the pre-look data are not well calibrated by the surface spline generated for the
remaining data. We assume this is due to changes in the CCD in the time that occurred between
the pre-look and the full set of observations. We also discard eclipses where less than half the event
is observed, one for TrES-2, one for WASP-3 and one for HAT-P-4. As discussed in Section 2.2,
we finally discard four of the nine WASP-3 secondary eclipses that fall on regions of the CCD we
cannot calibrate.
We fit the eclipses using a transit model with the best-fit parameters from the transit analysis
and no limb-darkening. We then scale the depth of this model to fit the data, finding the depth
that minimizes the χ2 value. For each target we then find the mean (¯x) and standard deviation
(σx) of the individual best-fit depths, and define the 95% confidence upper limit on the eclipse
depth as ¯x + 2σx. The secondary eclipses of HAT-P-4, TrES-3, TrES-2 and WASP-3 are shown in
-- 19 --
Fig. 8. -- Upper panel: The nine TrES-2 transits, phase-folded and binned in two minute intervals.
The solid line is the best-fit transit model. Lower panel: The residuals when the best-fit model is
subtracted from the data.
Figures 9 -- 12. The upper limits are given in Table 6. We note that we achieve a useful constraint
only in the case of TrES-3.
5. Discussion
5.1. HAT-P-4
For HAT-P-4, our estimates of the system parameters are consistent with (and in the case of
inclination, more precise than) those published by Kovacs et al. (2007) and Torres et al. (2008).
We calculate Rp = 1.332 ± 0.052 RJup, R⋆ = 1.602 ± 0.061 R⊙, i = 89.67 ± 0.30 degrees and
τ = 255.6 ± 1.9 minutes, where τ is the transit duration from first to fourth contact. We use the
discovery epoch and the ten EPOCh transit times presented in this paper to produce a new refined
ephemeris of Tc(BJD) = 2454245.81531 ± 0.00021 + 3.0565114 ± 0.0000028E. Figure 14 shows the
residuals to the new ephemeris. We see no evidence for transit timing variations in the residuals
which have a scatter of roughly 2 minutes.
-- 20 --
Fig. 9. -- Upper panel: The eight WASP-3 transits, phase-folded and binned in two minute intervals.
The solid line is the best-fit transit model. Lower panel: The residuals when the best-fit model
is subtracted from the data. The significant in-transit deviation from the model is discussed in
Section 5.
We use eight of the nine observed secondary eclipses to constrain the depth of the eclipse in
the EPOCh bandpass, discarding the ninth due to poor coverage of the event. The eclipses are
shown in Figure 10. We set a 95% confidence upper limit on the eclipse depth of 0.073%, which, if
it were produced entirely by reflected light, would correspond to a planetary geometric albedo of
Ag = 3.5, a physically impossible value. In the future, full phase curves of HAT-P-4 are scheduled
to be observed in the near-infrared 3.6 and 4.5 micron IRAC bands, as part of the Warm Spitzer
census of exoplanet atmospheres, at which point we may begin to study the atmosphere in more
detail.
5.2. TrES-3
For TrES-3, we find system parameters consistent with those published by O'Donovan et al.
(2007), Sozzetti et al. (2009) and Gibson et al. (2009), with Rp = 1.320± 0.057 RJup, R⋆ = 0.817±
-- 21 --
Fig. 10. -- Eight EPOCh secondary eclipse observations of HAT-P-4, offset in relative flux for
clarity and binned in five minute intervals. The error on each point is σ/√N, where σ is the scatter
in the bin and N the number of points. The solid lines are the best fit eclipse model in each case.
The bottom three eclipses were obtained in the contingent block of observations.
0.022 R⊙, i = 81.99 ± 0.30 degrees and τ = 81.9 ± 1.1 minutes. In the upper panel of Figure 15
we plot the published values of inclination with time and note a weak trend towards decreasing
inclination, however it is present at only the 1.5σ level, and hence not significant (and largely
dependent on the most recent value from Sozzetti et al. 2009). A more model-independent way
of constraining changes in the transit parameters with time is by measuring the transit duration.
Where available, we use the quoted transit duration and error, and otherwise we calculate the
Table 6. EPOCh secondary eclipse measurements
Target
Eclipse Depth
Upper limit
Implied Ag
HAT-P-4 −0.0069 ± 0.0397%
TrES-3
TrES-2
WASP-3
−0.020 ± 0.041%
0.023 ± 0.071%
0.023 ± 0.044%
0.073%
0.062%
0.16%
0.11%
3.5
0.81
6.6
2.5
-- 22 --
Fig. 11. -- Six EPOCh secondary eclipse observations of TrES-3, offset in relative flux for clarity.
The error on each point is σ/√N , where σ is the scatter in the bin and N the number of points.
The solid lines are the best fit eclipse model in each case.
transit duration from the published parameters, using equation (4) from Charbonneau et al. (2006).
Following the analytic approximation of Carter et al. (2008), we set the error on these calculated
transit durations to twice the error in the measured transit times for each source. Although this
error was originally derived for the transit duration from mid-ingress to mid-egress, as compared
to the transit duration from first to fourth contact, we find that for the EPOCh data the errors
calculated using this approximation and the errors measured from the data themselves are nearly
identical (1.1 minutes and 1.0 minutes respectively). We plot the derived values in the lower panel
of Figure 15, and we see no evidence of a change in the transit duration with time.
In Section 2.2 we noted that in the process of calibrating the light curve, a long term variability
was evident. This variability is consistent with stellar variability due to spots. Using a vsini of <2
km s−1 (O'Donovan et al. 2007), the rotational period of TrES-3 must be >21 days, considerably
longer than our observation span of 12 days. We can therefore not place any additional constraints
on the rotational period of TrES-3, however we note that if the variability is due to spots on the
stellar surface rotating in and out of view then additional monitoring of TrES-3 may reveal the
rotational period.
-- 23 --
Fig. 12. -- Six EPOCh secondary eclipse observations of TrES-2, offset in relative flux for clarity.
The error on each point is σ/√N , where σ is the scatter in the bin and N the number of points.
The solid lines are the best fit eclipse model in each case.
For TrES-3, we calculate a new ephemeris of Tc(BJD) = 2454538.58069±0.00021+1.30618606±
0.00000038E using the published transit times and the seven EPOCh transits presented in this
paper. Figure 16 shows the residuals to the new ephemeris. We see no evidence of the period
changing with time or transit timing variations larger than 1 minute.
Using the six EPOCh secondary eclipse observations of TrES-3, shown in Figure 11, we set a
95% confidence upper limit on the eclipse depth of 0.062%. This indicates the planetary geometric
albedo must be Ag < 0.81 in the EPOCh bandpass. Winn et al. (2008) observed secondary eclipses
of TrES-3 in the i, z and R bands, and were able to put 99% confidence upper limits on the
eclipse depths of 0.024%, 0.050% and 0.086% respectively. The EPOCh upper limit at 0.65 µm is
consistent with the R band upper limit. de Mooij & Snellen (2009) observed the secondary eclipse
in the K band and found a depth of 0.241±0.043%. Fressin et al. (2010) observed secondary eclipses
of TrES-3 with the Spitzer IRAC instrument, measuring depths of 0.356±0.036%, 0.372±0.054%,
0.449±0.097% and 0.475±0.046% in the 3.6, 4.5, 5.8 and 8.0 micron bands respectively. The
secondary eclipse measurements are shown in Figure 17 as a function of wavelength.
-- 24 --
Fig. 13. -- Four EPOCh secondary eclipse observations of WASP-3, offset in relative flux for clarity.
The error on each point is σ/√N , where σ is the scatter in the bin and N the number of points.
The solid lines are the best fit eclipse model in each case.
Given the high levels of stellar irradiation, the atmosphere of TrES-3 was anticipated to host
a thermal inversion (Fortney et al. 2008; de Mooij & Snellen 2009). Using all data sets, however,
Fressin et al. (2010) found the observations to be best fit with a dayside atmosphere model without
a thermal inversion.
Our model spectra are computed using the exoplanet atmosphere model developed in Madhusudhan & Seager
(2009). The model consists of a line-by-line radiative transfer model, with constraints of hydrostatic
equilibrium and global energy balance, and coupled to a parametric pressure-temperature (P-T)
structure and parametric molecular abundances (parametrized as deviations from thermochemical
equilibrium and solar abundances). Our modeling approach allows one to compute large ensembles
of models, and efficiently explore the parameter space of molecular compositions and temperature
structure.
We confirm previous findings that existing detections of day-side observations can be explained
to within the 1σ uncertainties by models without thermal inversions. The black curve in Figure 17
shows one such model spectrum, which has a chemical composition at thermochemical equilibrium
-- 25 --
Fig. 14. -- The transit times of HAT-P-4. The open diamonds are the EPOCh transit times from
this paper; the asterisk is the discovery epoch (Kovacs et al. 2007). Lower panel: An expanded
view of the EPOCh transit times, with 1σ errors of 48-71 seconds.
and solar abundances for the elements. The model is also consistent with the EPOCh upper-limit
at 0.65 microns, and with the upper-limits from Winn et al. (2008). The dark green dashed curve
shows a 1600K blackbody spectrum of the planet, indicating that the data cannot be explained
by a pure blackbody. The model reported here has a day-night energy redistribution fraction of
0.4, indicating very efficient redistribution. Therefore, based on previous studies and our current
finding, existing data do not require the presence of a thermal inversion in TrES-3. However, a
detailed exploration of the model parameter space would be needed to rule out thermal inversions
with a given statistical significance (Madhusudhan & Seager 2010).
5.3. TrES-2
For TrES-2, we derive system parameters that are consistent at the 1.5σ level with estimates
published by O'Donovan et al. (2006), Sozzetti et al. (2007), and Holman et al. (2007), finding
Rp = 1.169 ± 0.034 RJup, R⋆ = 0.940 ± 0.026 R⊙, i = 84.15 ± 0.16 degrees and τ = 107.3 ± 1.1
-- 26 --
Fig. 15. -- Upper panel: The estimates of the inclination for TrES-3 as a function of time. Lower
panel: The estimates of the TrES-3 transit durations.
minutes.
As discussed in Section 1, there is currently a debate as to whether the inclination of the
planetary orbit and duration of the TrES-2 transit are decreasing with time due to orbital precession.
In the upper panel of Figure 18 we plot the estimates for the inclination as a function of time. For the
inclination, the error bars of Mislis & Schmitt (2009), Mislis et al. (2010) and Scuderi et al. (2009)
were calculated by fixing the stellar and planetary radii and allowing only the inclination and time
of center of transit to vary. The remainder of the inclination error bars were calculated allowing
all of the geometric parameters to vary simultaneously, which explains why they are considerably
larger than the later results. Since the errors skew any weighted linear fit towards an unrealistically
large increase in the inclination with time, we instead plot an unweighted linear fit to guide the
eye. We note that TrES-2 is in the Kepler field and that any change in inclination with time will
soon be measured with exquisite precision.
The inclination measured from a particular transit light curve will necessarily depend on the
geometric parameters and to some extent the choice of limb-darkening treatment. However, the
transit duration is directly measurable from the light curve and should not depend on the limb
-- 27 --
Fig. 16. -- The times of transit of TrES-3. O'D07: O'Donovan et al. (2007); S09: Sozzetti et al.
(2009); G09: Gibson et al. (2009); EPOXI: this paper. Lower panel: An expanded view of the
EPOCh transit times, with 1σ errors of 28 -- 35 seconds.
darkening. The lower panel of Figure 18 shows the published transit durations as a function of
time. Where they were not given, we calculated the durations and errors as described for TrES-3.
In this case, we perform a weighted linear fit and do see a formally significant decrease in the
transit duration with time. However, this conclusion is heavily dependent on one point, in this
case the duration calculated from Holman et al. (2007). If this point is excluded from the fit, then
dτ /dt = −0.0015±0.0015, consistent with no change in the transit duration with time and therefore
we do not claim to have detected a change in the transit duration with time. Again, we expect
Kepler to provide a clear answer to this question.
Using the published transit times of TrES-2 and the nine transits observed by EPOCh pre-
sented in this paper, we find a new weighted ephemeris of Tc(BJD) = 24544664.23039 ± 0.00018 +
2.47061344± 0.00000075E. The residuals to this ephemeris are shown in Figure 19. In the EPOCh
residuals, we see no variations in the transit times above the level of 2 minutes; excluding the
amateur data from the Exoplanet Transit Database due to the large error bars, the scatter in the
full set of residuals is less than 5 minutes. We see no evidence for long term drifts in the period.
-- 28 --
Fig. 17. -- The optical and near-infrared secondary eclipse measurements of TrES-3. The EPOCh
upper limit of 0.062% is shown in blue at 0.65 microns. The remaining upper limits in the optical
are from Winn et al. (2008); the measurement at 2.2 microns is from de Mooij & Snellen (2009);
and the four measurements from 3.6 to 8.0 microns are from Fressin et al. (2010). The solid black
line is a representative model from the set of models that fit the data to within 1σ, and the dashed
lined shows a black-body spectrum for a temperature of 1600K. The green circles represent the
model integrated to the Spitzer bandpasses. The inset is the temperature-pressure profile for the
model shown.
We used six of the eight EPOCh secondary eclipses of TrES-2 to place a 95% confidence upper
limit on the eclipse depth of 0.16%. This corresponds to a planetary geometric albedo of Ag = 6.6.
As for HAT-P-4, this is not a physically plausible value.
5.4. WASP-3
For WASP-3, we measure system parameters that are consistent with, and an improvement
upon, previously published parameters from Pollacco et al. (2008) and Gibson et al. (2008), finding
-- 29 --
Fig. 18. -- Upper panel: The estimates of the inclination of TrES-2 as a function of time. The
dotted line is an unweighted linear fit. O'D07: (O'Donovan et al. 2007); H07: (Holman et al. 2007);
R09: (Rabus et al. 2009); M09/10: (Mislis & Schmitt 2009; Mislis et al. 2010); S09: (Scuderi et al.
2009); EPOXI: this paper. Lower panel: The TrES-2 transit durations. In this case the dotted line
is a weighted linear fit.
Rp = 1.385±0.060 RJup, R⋆ = 1.354±0.056 R⊙, i = 84.22±0.81 degrees and τ = 167.3±1.3 minutes.
We generate a new refined ephemeris from the published transit times and the eight EPOCh transits
in this paper, finding Tc(BJD) = 2454686.82069± 0.00039 + 1.8468373± 0.0000014E. The residuals
to this ephemeris are shown in Figure 20.
The phase-folded light curve of WASP-3 (Figure 9) shows correlated residuals in the latter half
of transit. Since the noise in the transit exceeds the noise out of transit, one conclusion could be
spot activity on the surface of the star being eclipsed during transit. However, if we examine the
transits individually we observe that the correlated noise in the full light curve is not typically larger
in transit than out of transit. The six transits used in the analysis are shown in Figure 21. In the
transits numbered 2, 3, 4, and 6 large deviations can be seen in the second half of the transit, which
leads to residuals in the phased light curve. If there were star spots producing correlated residuals
in the transits, we would not necessarily expect them to occur at the same phase for each transit.
The v sin i for WASP-3 has been measured by Simpson et al. (2009) to be 15.7+1.4
−1.3 km s−1, which
-- 30 --
Fig. 19. -- Upper panel: The transit times of TrES-2. O'D06: O'Donovan et al. (2006); H07:
Holman et al. (2007); R09: Raetz et al. (2009); R09(ETD): Raetz et al. (2009) (from the Exoplanet
Transit Database, http://var.astro.cz/ETD); EPOXI: this paper. Lower panel: An expanded view
of the EPOCh transit times, with 1σ errors of 24 -- 44 seconds.
corresponds to a rotational period for the star of 4.2 days. Transits of WASP-3 are spaced by 1.85
days, so it is improbable for spot activity to appear at the same phase in successive transits. Given
these constraints, we conclude that the alignment of signals with phase in the EPOCh transits of
WASP-3 are coincidental and are due to instrumental artifacts.
We use four of the nine EPOCh secondary eclipse observations of WASP-3 to set a 95%
confidence upper limit on the eclipse depth of 0.11%. This corresponds to a planetary geometric
albedo of Ag = 2.5 in the EPOCh bandpass, which is not a useful constraint for the planetary
atmosphere.
6. Conclusion
We have presented time series photometry from the NASA EPOXI Mission of Opportunity
for four known transiting planet systems: HAT-P-4, TrES-3, TrES-2 and WASP-3. For each
-- 31 --
Fig. 20. -- The transit times of WASP-3. P08: Pollacco et al. (2008); G08: Gibson et al. (2008);
EPOXI: this paper. Lower panel: An expanded view of the EPOXI transit times, with 1σ errors
of 23 -- 51 seconds.
system we provided an updated set of system parameters and orbital period, and placed upper
limits on the secondary eclipse depth. For TrES-3, we see evidence of stellar variability over long
timescales. We combined the EPOCh secondary eclipse upper limit for TrES-3 with previously
published measurements and confirm that the data are best fit using an atmosphere model with
no temperature inversion. For TrES-2, the EPOCh data weaken the claimed trends of decreasing
inclination and transit duration (Mislis & Schmitt 2009; Mislis et al. 2010). We have also performed
a search for additional transiting planets in the EPOCh photometry for these systems, which we
will present in a forthcoming paper.
We are extremely grateful to the EPOXI Flight and Spacecraft Teams that made these difficult
observations possible. At the Jet Propulsion Laboratory, the Flight Team has included M. Abra-
hamson, B. Abu-Ata, A.-R. Behrozi, S. Bhaskaran, W. Blume, M. Carmichael, S. Collins, J. Diehl,
T. Duxbury, K. Ellers, J. Fleener, K. Fong, A. Hewitt, D. Isla, J. Jai, B. Kennedy, K. Klassen, G.
LaBorde, T. Larson, Y. Lee, T. Lungu, N. Mainland, E. Martinez, L. Montanez, P. Morgan, R.
Mukai, A. Nakata, J. Neelon, W. Owen, J. Pinner, G. Razo Jr., R. Rieber, K. Rockwell, A. Romero,
-- 32 --
Fig. 21. -- The six EPOCh transits of WASP-3. The best fit model for the combined set of transits
is plotted in each case. The scatter around the model is not typically larger in transit than out of
transit, indicating that the in-transit residuals cannot be attributed to star spots.
B. Semenov, R. Sharrow, B. Smith, R. Smith, L. Su, P. Tay, J. Taylor, R. Torres, B. Toyoshima,
H. Uffelman, G. Vernon, T. Wahl, V. Wang, S. Waydo, R. Wing, S. Wissler, G. Yang, K. Yetter,
and S. Zadourian. At Ball Aerospace, the Spacecraft Systems Team has included L. Andreozzi,
T. Bank, T. Golden, H. Hallowell, M. Huisjen, R. Lapthorne, T. Quigley, T. Ryan, C. Schira, E.
Sholes, J. Valdez, and A. Walsh.
Support for this work was provided by the EPOXI Project of the National Aeronautics and
Space Administration's Discovery Program via funding to the Goddard Space Flight Center, and
to Harvard University via Co-operative Agreement NNX08AB64A, and to the Smithsonian Astro-
physical Observatory via Co-operative Agreement NNX08AD05A. The authors acknowledge and
are grateful for the use of publicly available transit modeling routines by Eric Agol and Kaisey
Mandel, and also the Levenberg-Marquardt least-squares minimization routine MPFITFUN by
Craig Markwardt. This work has used data obtained by various observers collect in the Exoplanet
Transit Database, http://var.astro.cz/ETD.
-- 33 --
REFERENCES
Ballard, S., et al. 2010, ApJ, submitted, arXiv:0909.2875
Borucki, W. J., et al. 2009, Science, 325, 709
Carter, J. A., Yee, J. C., Eastman, J., Gaudi, B. S., & Winn, J. N. 2008, ApJ, 689, 499
Charbonneau, D., et al. 2006, ApJ, 636, 445
Christiansen, J. L., et al. 2010, ApJ, 710, 97
Claret, A., 2000, A&A, 363, 1081
de Mooij, E. J. W., & Snellen, I. A. G. 2009, A&A, 493, 35
Fortney, J. J., Lodders, K., Marley, M. S., & Freedman, R. S. 2008, ApJ, 678, 1419
Fressin, F., Knutson, H. A., Charbonneau, D., O'Donovan, F. T., Burrows, A., Deming, L. D.,
Mandushev, G., & Spiegel, D. 2010, ApJ, 711, 374
Gibson, N. P., et al. 2008, A&A, 492, 603
Gibson, N. P., et al. 2009, ApJ, 700, 1078
Hampton, D. L., Baer, J. W., Huisjen, M. A., Varner, C. C., Delamere, A., Wellnitz, D. D.,
A'Hearn, M. F., & Klaasen, K. P. 2005, Space Science Reviews, 117, 43
Holman, M. J., et al. 2007, ApJ, 664, 1185
Klaasen, K. P., Carcich, B., Carcich, G., Grayzeck, E. J., & McLaughlin, S. 2005, Space Science
Reviews, 117, 335
Kovacs, G., et al. 2007, ApJ, 670, 41
Kurucz, R. 1994, Solar abundance model atmospheres for 0,1,2,4,8 km/s. Kurucz CD-ROM
No. 19. Cambridge, Mass.: Smithsonian Astrophysical Observatory, 1994., 19
Kurucz, R. L. 2005, Memorie della Societa Astronomica Italiana Supplement, 8, 14
Madhusudhan, N., & Seager, S. 2009, ApJ, 707, 24
Madhusudhan, N., & Seager, S. 2010, ApJ, in press, arXiv:1010.4585
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Mislis, D., & Schmitt, J. H. M. M, 2009, A&A, 500, L45
Mislis, D, Schroter, S., Schmitt, J. H. M. M., Cordes, O., & Reif, K. 2007, A&A, 510, 107
-- 34 --
O'Donovan, F. T., et al. 2006, ApJ, 651, L61
O'Donovan, F. T., et al. 2007, ApJ, 663, L37
O'Donovan, F. T., Charbonneau, D., Harrington, J., Madhusudhan, N., Seager, S., Deming, L. D.
& Knutson, H. A. 2010, ApJ, 710, 1551
Pollacco, D., et al. 2008, MNRAS, 385, 1576
Raetz, St., et al. 2009, AN, 330, 459
Rabus, M, Deeg, H. J., Alonso, R., Belmonte, J. A., & Almenara, J. M. 2009, A&A, 508, 1011
Scuderi, L. J., Dittman, J. A., Males, J. R, Green, E. M., & Close, L. M. 2009, ApJ, submitted,
arXiv:0907.1685
Simpson, E. K., et al. 2009, MNRAS, submitted, arXiv:0912.3643
Sozzetti, A, et al. 2009, ApJ, 691, 1145
Sozzetti, A., Torres, G., Charbonneau, D., Latham, D., Holman, M. J., Winn, J. N., Laird, J. B.,
& O'Donovan, F. T. 2007, ApJ, 664, 1190
Torres, G., Winn, J. N., & Holman, M. J. 2007, ApJ, 677, 1324
Winn, J. N., et al. 2008, ApJ, 683, 1076
This preprint was prepared with the AAS LATEX macros v5.2.
|
1511.00952 | 1 | 1511 | 2015-11-03T15:46:22 | The iodine-plutonium-xenon age of the Moon-Earth system revisited | [
"astro-ph.EP"
] | From iodine-plutonium-xenon isotope systematics, we re-evaluate time constraints on the early evolution of the Earth-atmosphere system and, by inference, on the Moon-forming event. Two extinct radioactivites (129I, T1/2 = 15.6 Ma, and 244Pu, T1/2 = 80 Ma) have produced radiogenic 129Xe and fissiogenic 131-136Xe, respectively, within the Earth, which related isotope fingerprints are seen in the compositions of mantle and atmospheric Xe. Recent studies of Archean rocks suggest that xenon atoms have been lost from the Earth's atmosphere and isotopically fractionated during long periods of geological time, until at least the end of the Archean eon. Here we build a model that takes into account these results. Correction for Xe loss permits to compute new closure ages for the Earth's atmosphere that are in agreement with those computed for mantle Xe. The minimum Xe formation interval for the Earth- atmosphere is 40 (-10+20) Ma after start of solar system formation, which may also date the Moon-forming impact. | astro-ph.EP | astro-ph | Journal: Philosophical Transactions of the Royal Society A
The I-Pu-Xe age of the Moon-Earth system revisited
G. Avice1 & B. Marty1
1CRPG-CNRS, Université de Lorraine, 15 rue Notre-Dame des Pauvres, BP 20, F-54501 Vandoeuvre-
lès-Nancy Cedex, France
Author for correspondence: G. Avice, [email protected]
Cite this article: Avice G, Marty B. 2014 The iodine-plutonium-xenon age of the Moon-
Earth system revisited. Phil. Trans. R. Soc. A 372:20130260
Summary
From iodine-plutonium-xenon isotope systematics, we re-evaluate time constraints on
the early evolution of the Earth-atmosphere system and, by inference, on the Moon-forming
event. Two extinct radioactivites (129I, T1/2 = 15.6 Ma, and 244Pu, T1/2 = 80 Ma) have produced
radiogenic 129Xe and fissiogenic 131-136Xe, respectively, within the Earth, which related
isotope fingerprints are seen in the compositions of mantle and atmospheric Xe. Recent
studies of Archean rocks suggest that xenon atoms have been lost from the Earth's atmosphere
and isotopically fractionated during long periods of geological time, until at least the end of
the Archean eon. Here we build a model that takes into account these results. Correction for
Xe loss permits to compute new closure ages for the Earth's atmosphere that are in agreement
with those computed for mantle Xe. The minimum Xe formation interval for the Earth-
+20 Ma after start of solar system formation, which may also date the
atmosphere is 40-10
Moon-forming impact.
Keywords: Moon, Xenon, age, atmosphere
© 2015. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
1. Introduction
The age of the solar system is well established at 4.568 Ga [1–3]. Extant and extinct
radioactivity systems indicate that not only primitive bodies, but also differentiated
planetesimals and planetary embryos including Mars formed within a few Ma after start of
condensation in the solar system (inferred from the age of calcium-aluminium rich inclusions
- CAIs - in primitive meteorites). In contrast, the formation age of the Earth-Moon system is
uncertain and is presently debated within a time interval of 30-200 Ma after CAI (this
volume). Deciphering details of the early chronology of the Earth requires the development of
adequate extinct radioactivity chronometers. Because the Earth's interior has been well mixed
by mantle convection over 4.5 Ga, most of the early reservoirs have been re-homogenized
even if some remnants of past heterogeneities might be still present [4–7]. However,
information on ancient reservoirs is still kept at the Earth's surface, in old terranes, and, in the
case of noble gases, in the terrestrial atmosphere.
Xenon, the heaviest noble gas, has a large number (9) of isotopes, and extant and extinct
radioactivity products have contributed several of them. Iodine-129 decays with a half-life of
15.7 Ma into 129Xe [8], resulting in 129Xe excesses in primitives meteorites relative to the
potential primordial Xe isotopic compositions [9]. Atmospheric Xe presents a monoisotopic
excess of 129Xe (compared to adjacent 128Xe and 130Xe isotopes) of about 7% (e.g. [10])
attributed to the decay of extinct 129I. Some natural gases and mantle-derived rocks [11–13]
have 129Xe/130Xe ratios (where 130Xe is a stable isotope of xenon which is used for
normalization) higher than the atmospheric value. Altogether, these observations demonstrate
that the Earth formed and differentiated while 129I was still present, thus within a few tens of
Ma. Most (>80 %) of terrestrial Xe is now in the atmosphere (e.g. [11] but see [14] for an
alternate view), so that atmospheric Xe is to a first order representative of total terrestrial Xe.
Consequently, a 129I-129Xe age of the Earth can be constrained from estimates of the initial
abundance of iodine, inferred from the present-day abundance of the stable isotope 127I [15].
Although the latter is not well known (probably not better than a factor of 2, see below), the
exponential nature of radioactive decay makes the result not so sensitive to this uncertainty.
Thus the I-Xe age of the Earth's atmosphere, which is in fact the time interval Δt129 of
reservoir closure, can be expressed as:
∆𝑡!"#= !!!"#ln (
!!"!
!"#!"!
!"#
)
(1)
2
where λ129 is the decay constant of 129I (4.41x10-2 Ma-1), 129IINI is the initial 129I abundance in
the reservoir. The latter is computed with the (129I/127I)INI initial ratio from meteorite data
(1.1x10-4) [16] and estimates of terrestrial 127I abundance (>3 ppb, up to 13 ppb, see next
section). 129Xe(I) represents the 129Xe excess resulting from the decay of 129I in the
atmosphere (2.8x1011 mol. of 129Xe(I) [10]). Within these assumptions, the Earth would have
become closed for Xe isotope loss at 100-120 Ma after CAI, this range depending mostly on
the initial abundance of iodine (for further discussion of these parameters, see reviews by
[10,15,17–19]). This is the classical "age" of the atmosphere found in textbooks.
The other short-lived nuclide of interest here is 244Pu (half-life of 80 Ma [20]) which, in
addition to α-decay, presents a weak (0.125 %) branch for spontaneous fission and produces
131,132,134,136Xe isotopes (represented hereafter by 136Xe(Pu)). These heavy Xe isotopes are also
produced in Earth by the spontaneous fission of extant 238U. However the contribution of 238U
fission to Xe isotopes was minor compared to that of 244Pu during the periods of time
characterizing the Earth's formation and its early evolution. Contrary to iodine, plutonium has
no stable isotope, so that the initial abundance of 244Pu is inferred from comparison with U in
meteorites and Earth [21], since both are refractory and lithophile elements. Fissiogenic Xe
from 244Pu has been found in the Earth's interior [4,12,22]. The detection of fissiogenic Xe in
the atmosphere is, however, not straightforward, as the original composition of atmospheric
xenon is not directly measurable. In fact, atmospheric xenon is isotopically fractionated by 3-
4 % per atomic mass unit relative to potential primordial candidates [23]. Furthermore, even
after correction for such mass-dependent isotope fractionation, neither chondritic nor solar Xe
can be directly related to atmospheric Xe, because both chondritic and solar Xe are rich in the
heavy Xe isotopes (134Xe and 136Xe) compared to "unfractionated" atmospheric Xe [24].
Takaoka [25], and Pepin and Phinney [26] extrapolated, from meteorite data, a primordial Xe
component (labelled Xe-U by [26]), from which atmospheric Xe could be derived by mass-
dependent isotopic fractionation. Xe-U has an isotope composition close to solar Xe for its
light masses but is depleted in 134Xe and 136Xe. This U-Xe component has still not been found
in meteorites, possibly because of
the presence of superimposed components of
nucleosynthetic origin. Indeed xenon trapped in different meteoritic phases presents variations
in its s-, p- and r- process isotopes (e.g., the P3, P6 and HL components trapped in
nanodiamonds [27,28]). Thus the heavy Xe isotope difference between potential Xe ancestors
and solar Xe could be the result of different mixes of nucleosynthetic Xe isotopes in primitive
reservoirs, with the possibility that solar Xe was more contributed by s-process isotopes
3
compared to other Xe primordial progenitors [29]. Whatever the composition of the
progenitor of atmospheric Xe, this reservoir appears poor in 244Pu-derived Xe isotopes.
Estimates of 136Xe excess in the atmosphere due to contribution of extinct 244Pu vary
between 4.6 % and 2.8 %, according to Pepin and Phinney [26] and Igarashi [30],
respectively. 136Xe(Pu) gives another possibility to estimate closure ages for the Earth-
atmosphere. Both the I-Xe and Pu-Xe systems can be combined yielding an I-Pu-Xe time
interval Δt129-244 of: ∆𝑡!"#!!""=
!!""!!!"#𝑙𝑛
!
!"(!)
!"#
!"(!")
!"#
! !"#
!" !""
!"!
𝑌!""
!"#
(2)
where λ244 is the decay constant of 244Pu (8.45x10-3 Ma-1), and 136Y244 is the production yield
of 136Xe from 244Pu fission (7x10-5, [31]). The 129Xe(I)/136Xe(Pu) ratio of the atmosphere has
been estimated to be 4.6 [32], which yields an atmospheric Δt129-244 closure time of about 100
Ma after CAI, consistent with the I-Xe age. The fact that these two closure ages are
comparable is not merely a coincidence since Δt129-244 depends in large part on the residual
amount of 129Xe(I) in Earth's atmosphere (of the order of 1 %) rather than on that of 136Xe(Pu)
in this time interval, given the much shorter half-life of 129I compared to that of 244Pu. A more
interesting constraint arises form a direct comparison of the amount of 136Xe(Pu) left in Earth
(atmosphere) to that potentially produced by initial 244Pu. Although the latter has also a
significant uncertainty, it appears that a large fraction of 136Xe(Pu) (≥70 %) is missing in the
present-day Earth's atmosphere [17,32–34]. Given the half-life of 244Pu of 80 Ma, this
discrepancy suggests that Xe was lost from the atmosphere after the giant impact phase of the
Earth's accretion [17,33].
Atmospheric xenon is not only isotopically fractionated (enriched in heavy isotopes by
3-4 % per atomic mass unit) compared to potential primordial Xe, but is also elementally
depleted by one order of magnitude relative to other noble gases (e.g., Kr) compared to the
abundance pattern of meteoritic noble gases. These dual characteristics, known for long as the
"xenon paradox" (elementally depleted in heavy element, isotopically enriched in heavy
isotopes) have not yet found a satisfactory explanation. The Xe depletion and the lack of Pu-
produced Xe isotopes in the atmosphere suggest that, after a last giant impact event, there
might have been more xenon in the atmosphere, which would have been lost through a
4
process that fractionated Xe isotopes (if both features were related). Therefore the I-Pu-Xe
ages should be corrected for Xe loss and the scope of the correction would depend on the
timing of loss relative to I and Pu decays. The apparent deficiency of 136Xe(Pu) in the
atmosphere (see previous paragraph) may indeed be a consequence of prolonged selective
loss of atmospheric Xe.
Recent studies of noble gases in Archean (3.5-3.0 Ga-old) rocks may provide a solution
to the xenon paradox. Isotopically fractionated Xe has been found in Archean barite [35,36]
and hydrothermal quartz [37,38]. The Xe isotopic spectrum is intermediate between the
primordial and the modern atmospheric Xe isotope patterns, and the isotopic fractionation
(relative to the modern composition) tends to decrease with decreasing age (Fig. 1). Together
with Xe data from ancient basement fluids of presumed Proterozoic age [39], the evolution of
Xe isotopic fractionation with time is consistent with a Rayleigh distillation in which Xe has
been lost from the atmosphere with an instantaneous fractionation factor of about 1.1 % per
atomic mass unit [38]. The magnitude of the latter is in agreement with experimental studies
of Xe isotope fractionation upon ionization [40]. The exponential decrease of Xe isotope
fractionation with time is qualitatively consistent with that of the far UV light (FUV) flux
from the evolving Sun with time ([41]; Fig. 1) suggesting that Xe was selectively ionized and
lost from the atmosphere to space through time [42,43] at a rate that followed the declining
FUV flux.
In this study, we investigate the possibility to reproduce the current features of
atmospheric Xe (elemental and isotopic compositions), taking into account this long-term
escape. We develop a 3-box model (solid Earth, atmosphere, space) that allows us to correct
the abundances of radiogenic/fissiogenic Xe isotopes for Xe loss. Previous computed ages are
therefore not valid anymore and their values and meaning have to be revisited. Doing so, we
follow Podosek & Ozima [17] who predicted that "If allowance is made for the possibility
that most of the Xe, including radiogenic Xe, that should be in the atmosphere, has somehow
be removed or hidden, the I-Xe and Pu-Xe formation interval could be reduced to perhaps
60 Ma". We now have observational evidence for such Xe loss through geological time. Since
the atmosphere is probably very sensitive to impact-driven erosion [44,45], corrected closure
ages may be related to the end of the giant impact epoch that led to the formation of the Moon
[46].
5
2. Building of the model
The model consists of three reservoirs: the silicate Earth, the atmosphere and the outer
space (Fig. 2). We aim to estimate the closure time of the atmosphere Δt, defined here at the
time after CAI when the atmosphere became closed to volatile loss (except for Xe
preferentially lost during the Hadean and the Archean eons). Between time 0 (CAI) and Δt,
volatile elements are contributed to the proto-Earth by accreting bodies, and partially lost
through collision and atmospheric erosion. Between Δt and Present, only Xe is lost to space,
the other volatile elements being conservative in the atmosphere. We correct for this
secondary Xe loss using the depletion of xenon relative to other noble gases in the
atmosphere.
Xenon is degassed without isotopic fractionation from the Earth's interior to the
atmosphere through magmatism. Between Δt and the end of the Archean eon, xenon escapes
from the atmosphere to the outer space, and is isotopically fractionated during this escape.
Three radioactive systems are involved: 129Xe produced by the β-decay of 129I (T1/2=15.6 Ma),
and 131,132,134,136Xe from the fission of 244Pu (T1/2=80 Ma) and 238U (T1/2=4.47 Ga). As 136Xe,
compared to other xenon isotopes (131,132,134Xe), is a major product from the fission of 244Pu, it
will be considered as a proxy for the entire fission component in the following discussion.
The following mass balance exemplifies the evolution of the atmospheric 129XeATM
(mol.) with time:
!
!"!"#(!)
!"# !"
=𝜑𝑡
𝑋𝑒!"#$
!"#
𝑡 −𝛽(𝑡)(1+𝛼!"#) 𝑋𝑒!"#(𝑡)
!"#
(3)
where 129XeATM(t) is the abundance of 129Xe in the atmosphere at time t. 129XeMANT(t)
represents the abundance of 129Xe atoms in the mantle at time t. φ(t) and β(t) are the degassing
and escape parameters, respectively. αesc is the isotopic fractionation factor described below
(Eqn. 6). The evolution of the mantle 129XeMANT (mol.) with time is expressed by:
!
!"!"#$(!)
!"# !"
=−𝜑𝑡
𝑋𝑒!"#$
!"#
𝑡 +𝜆!"#
𝐼 !"#(𝑡)
(4)
where φ(t) is the degassing parameter at time t, λ129 is the decay constant of 129I into 129Xe,
129I(t) the abundance of iodine-129 in the mantle at the time t. Similarly, equations for
6
124-136Xe are defined taking into account the decays of the different radioactive nuclides
(244Pu, 238U). For example, for 136XeMANT, the equation is:
!
!"!"#$(!)
!"# !"
=−𝜑𝑡
𝑋𝑒!"#$
!"#
+𝜆!"#𝐵!"#
𝑡 +𝜆!""𝐵!""
!"#
𝑈 !"# (𝑡)
𝑌!"#
!"#
𝑌!""
𝑃𝑢 !""
𝑡 (5)
where 136XeMANT(t) is the abundance of 136Xe in the mantle at time t, 244Pu(t) and 238U(t) are
the abundances of parent nuclides in the mantle at time t, B244 (1.25x10-3) and B238
(5.45x10-7) are the branching ratios for 244Pu and 238U respectively, 136Y244 (6.3 %) and 136Y238
(5.6 %) are the yields of fission [31].
Equations are resolved with 1 Ma-step using an original code written with the
Mathematica® programming language. Results of the model comprise, for each temporal step,
the amount of each stable
the amount of each
radiogenic/fissiogenic isotope (e.g. 136Xe in the atmosphere coming from the fission of 244Pu).
in each reservoir plus
isotope
(a) Degassing from the Earth's interior
The rate of Xe degassing (φ(t) in equations (3-5)) from the Earth's interior through time
can be anticipated from thermal and geochemical considerations. Having a constant degassing
parameter φ(t) through Earth's history would result in a too severe Xe loss from the mantle.
We have tested different functions for φ(t) and model results that best fit the data are those
obtained using three different values (φ 1 > φ 2 > φ 3) for the respective intervals of time [Δt,
100 Ma], [100 Ma, 1000 Ma], [1000 Ma, 4500 Ma] (Table 1). An exponential decrease of φ(t)
could also fit the data but not as well as this stepped function.
The choices of these steps and of the related time intervals have some physical ground.
Due to a higher thermal regime of the solid Earth the degassing rate during the Hadean eon
was probably an order of magnitude higher than the modern one [18,47–49]. During the
Archean eon, the degassing rate was also probably higher than at present, as indicated for
example by the ubiquitous presence of komatiitic lavas, presumably originating from a mantle
hotter than today (although some authors argued that a hotter mantle does not necessarily
imply an higher convection rate [14,50]). Isotopic fractionation of xenon during magma
generation and degassing could only be kinetic, if any, and is neglected here. The model is
built in a way that the mantle is degassing Xe into the atmosphere from the time of Earth's
7
accretion, which is mathematically equivalent, during the short time interval of a few tens of
Ma, to add Xe from impacting bodies directly to the atmosphere.
(b) Loss of xenon to the outer space
As introduced above, Archean rocks of surficial origin contain mass-fractionated Xe
isotopically intermediate between Chondritic/Solar and modern Atmospheric (Fig. 2). Data
are scarce because they are extremely difficult to obtain (due to the need to date confidently
the host phases). Available data are consistent with a time evolution of the Xe isotopic
fractionation, presumably in the ancient atmosphere. This evolution can be fitted with a
Rayleigh distillation model in which Xe isotopes are escaping from the atmosphere through
time with mass-dependent instantaneous isotopic fractionation. The exponential evolution of
the isotopic composition of atmospheric xenon with time predicted by Rayleigh distillation is
consistent with the decline of FUV light flux from the ancient Sun (Fig. 1), suggesting that
interactions between atmospheric Xe and FUV light from the Sun played a role in Xe escape.
The Rayleigh distillation equation can be written as:
(
!"
!!!!" !
)!!!.!" !",!"#$%(
)!!!=𝑓!!"# !!~1.035
!"
!!!!" !
(6)
where the factor 1.035 relates to the isotopic difference between Solar/Chondritic and modern
air and f is the depletion factor of Xe in modern air corresponding to a factor of ≈20 relative
to carbonaceous chondrites [23]. The instantaneous fractionation factor αesc is then
≈1.011 % per atomic mass unit. This isotope fractionation is large for an inert gas with such a
high mass. Thus either Xe isotopic fractionation resulted from a specific process during
atmospheric escape, which is not yet documented, or it involved ionization of xenon, which,
from laboratory experiments, has been shown to yield isotopic fractionation of the order of
0.8-1.6 % per atomic mass unit [40,51–55]. Among most volatile species that were potentially
present in the ancient atmosphere (e.g., noble gases, CO, CO2, N2, CH4...), xenon has the
lowest ionization potential. Hébrard and Marty [42] have investigated the possible behaviour
of atmospheric Xe, taking into account the inferred distribution of FUV light wavelengths of
the ancient Sun. They proposed that Xe was ionized at an atmospheric height comparable to
that of organic haze formation in the Archean atmosphere, so that ionized heavy Xe isotopes
were preferentially retained in the lower atmosphere while ionized light Xe isotopes could
escape from the upper atmosphere. This possibility is certainly not unique and other processes
8
may be explored but, whatever the origin of this isotopically fractionating Xe loss, it remains
firm that Xe isotopically intermediate between primordial Xe and modern atmospheric Xe has
been found trapped in Archean rocks.
In our model, we use an instantaneous fractionation factor αesc allowed to vary within
1.0-1.5 % per atomic mass unit. The rate of escape of xenon atoms is scaled using FUV decay
curves corresponding to the wavelength at which xenon atoms are ionized (102.3 nm) [41].
The intensity of escape (β(t) in Ma-1) with time t is given by:
𝛽𝑡 =1𝑑(𝑏×𝑡)! (7)
b and c are constant parameters from [41] and d is an adjusted constant. Parameters values are
shown in Table 1 and the decay of atmospheric Xe with time is shown in Fig. 2.
3. Key parameters
The key parameters of the model are either taken directly from the literature (e.g.
(129I/127I)INI) when the value is widely accepted or are adjusted testing a range of values when
the value is badly known (e.g. IINI). Table 1 contains canonical values for parameters of the
model and adjusted values.
(a) Iodine content of the Earth
Because of its volatility, the precise iodine content of the Earth, and thus the 129IINI, is
not well known. A large range between 3 to 13 ppb is proposed in the literature [56–58]. The
lower value of 3 ppb is computed with iodine data from the depleted mantle [58] which, by
definition, is poor in incompatible elements like iodine. Thus it represents a lower limit of the
iodine budget of the Earth. Here we use a value of 6.4 ppb in our reference solution, which
gives model results consistent with observed Xe data, and permits a good match between the
two chronometers (I-Xe and I-Pu-Xe). The range of IINI values is used as an uncertainty and is
propagated in the age calculation. An abundance of 6.4 ppb for 127I together with the initial
129I/127I ratio of 1.1x10-4 obtained from meteorite data [16] yields an initial amount of
terrestrial 129I (129IINI) of 3.8x1012 mol.
(b) U-Pu content of the Earth
9
The range of present-day abundances for 238U is between 16 and 20 ppb [18]
corresponding to initial values, 4.57 Ga ago, between 33 and 41 ppb. Here we take a value of
38 ppb for the initial 238U content as in [18]. The average value of 6.8x10-3 for the
(244Pu/238U)INI is derived from the analysis of meteorites [59,60] and is consistent with data
obtained from ancient terrestrial zircons [18,61–63].
(c) Initial amount and isotopic composition of xenon
Xenon is under-abundant in the terrestrial atmosphere relative to other noble gases by a
factor of 23±5 relative to Kr [23]. Thus the present-day atmospheric inventory (6.15x1011
mol. of 130Xe [31]) can be corrected for selective escape of xenon atoms by multiplying the
current abundance by this factor, which leads to an initial 130Xe amount of 1.41(±0.36)x1013
mol. It is beyond the scope of this study to evaluate the starting isotopic composition of
primordial xenon and we refer to discussions in [17,26,32]. The U-Xe [23] is, so far, the only
initial isotopic composition permitting to reproduce the current isotopic composition of
atmospheric Xe including fissiogenic and radiogenic contributions [64] and its composition is
adopted as a starting isotopic composition for the model.
(d) Optimisation of the parameters
The model is constrained using the initial Xe amount of Earth and the U-Xe
composition on one hand, and the present-day amount and composition of atmospheric Xe
and, for some of the isotopic ratios, of mantle Xe on another hand. The silicate Earth contents
of parent elements (I and Pu) are additional parameters that can be adjusted within a plausible
range of values. The model must yield ancient Xe isotopic compositions that are consistent
with data from Archean rocks.
Key parameters from the literature are 130XeINI, (129I/127I)INI, 238UINI, (244Pu/238U)INI and
the starting isotopic composition (U-Xe). Other parameters are adjusted following multiple
runs. 130Xe, the stable isotope of reference, is used to scale the degassing and escape rates
over time. The different degassing rates φ1, φ2 and φ3 (Table 1) are fitted in order to respect
the 130Xe depletion over time and to take into account variations in the degassing rate of the
mantle. The φ1, φ2 and φ3 values that fit best the data are 890, 15 and 1 times the modern rate
between Δt and 100Ma, 100 Ma and 1 Ga, and 1 Ga and Present, respectively. The variation
of the atmospheric Xe escaping rate (Eqn. 7) is scaled to the variation of the FUV light flux
over time. Once the degassing and escaping rates of 130Xe are scaled, it is possible to optimize
10
other key parameters such as the instantaneous fractionation factor αesc and the initial iodine
abundance IINI. The aim of this final step of optimization is to reproduce the modern and the
Archean isotopic compositions of atmospheric Xe within better than 1% (Fig. 2).
(e) Outcomes of the model
In addition to optimisation of the degassing and escaping parameters, the model allows
one to determine the atmospheric closure time Δt that fits best the observations. Δt
corresponds to the time after CAI when the atmosphere became closed (except for Xe which
escape to space continued over eons).
4. Discussion
The long-term escape affects the global budget of xenon atoms in the Earth's
atmosphere and therefore the amount of radioactive products, which have to be corrected as
shown in Table 2.
(a) Corrected I-Xe age of the Earth-atmosphere
The I-Xe closure age of the atmosphere, corrected for long-term escape of atmospheric
Xe, becomes 41 Ma after CAI, instead of 110 Ma, with our reference I content of 6.4 ppb. For
a possible range of 3-13 ppb for terrestrial iodine [56–58], the range of closure age becomes
21-62 Ma (see ranges of solutions in Fig. 3). It must be noted that the I abundance of 3 ppb
[57] and the corresponding closure age of 21 Ma after CAI are likely to be a lower limit, so
that a ≈ 30-60 Ma possible range, corresponding to a terrestrial I content of 6-13 ppb, is
preferred.
(b) I-Pu-Xe closure age
The I-Xe closure age given above is in fact a mass balance between the residual amount
of radiogenic 129Xe in the atmosphere and the amount of initial 129I, with the underlying
assumption that the former derived from the latter. In fact, the whole amount of 129Xe(I) could
have been added to a "dead" Earth (that is, an Earth closed after the complete decay of 129I),
by an accreting primitive body formed just after CAIs and having therefore a high 129Xe(I)/I
ratio. In such an extreme case, the "closure" age based on I-Xe would not have any
chronological meaning for the Earth. Here it becomes interesting to make use of the I-Pu-Xe
system in addition to the I-Xe system. The (129Xe(I)/136Xe(Pu))ATM,CORR ratio depends on two
11
decay constants having contrasted values. Thus this ratio is time-dependent and decreases
with age. In the example given above, the (129Xe(I)/136Xe(Pu))ATM,CORR ratio should yield a
very young age, contrary to the I-Xe chronometer. In other words, the two chronometers all-
together have the faculty to give, or not, "concordant" closure ages.
The atmospheric 129Xe(I)/136Xe(Pu) ratio, corrected for Xe escape, becomes 21.7 instead
of 4.6 (Table 2), yielding a closure age of 34 Ma using Eqn. 2, with a possible range of 13-58
Ma. However, the lower time limit is based again on an unrealistically low terrestrial amount
of iodine. This range is "concordant" with the I-Xe range of 30-60 Ma and both methods
+20 Ma for the atmosphere, instead of about 100
converge to a revised closure age of ≈ 40-10
Ma previously thought (Fig. 3). Interestingly, the 129Xe(I)/136Xe(Pu) ratio has also been
estimated for the mantle from the precise Xe isotope analysis of mantle-derived CO2-rich
gases [32]. Estimates of this ratio vary between 20 and 64, depending on parameters (e.g.
I/Pu/U ratios, primordial Xe isotopic composition) chosen to derive them. This range of
values leads to a possible interval of 20-50 Ma for a major episode of mantle degassing [32].
Thus it is tempting to link such an event to a major catastrophic impact that affected the
interior of the proto-Earth and its atmosphere and that led to the formation of the Moon.
However, it must also be noted that Xe isotope data for mid-ocean ridge basalts seem to
define lower 129Xe(I)/136Xe(Pu) ratios around 6-10 [4,5], that could represent another giant
event [32], or, in our opinion, fractionation of volatile iodine with respect to refractory
plutonium/uranium for the mantle source of MORBs, during Earth's building episodes.
The range of closure ages (30-60 Ma after CAI) is consistent with Hf-W model ages for
the Moon's formation predicting an ancient fractionation (25-50 Ma) from the W isotope
difference between chondrites and Earth [65–68]. However, it is only marginally consistent
with those based on the similarity of the W isotopes ratios between Earth and Moon, that
requires the Moon formation event to have taken place ≥ 60 Ma after CAIs [69]. The former
case would reinforce the link between the last episode of major loss of the atmosphere and the
formation of the Moon.
(c) Physical significance of closure ages and the epoch of the formation of the Earth-Moon
system
The model presented here permits to correct the xenon budget of the Earth's atmosphere
for long-term escape to space. This correction leads to a reconstruction of the xenon isotopic
12
composition of the atmosphere just after the last episode of major loss. After correction, the I-
Xe and the I-Pu-Xe systems converge towards a common range of closure ages within 30-60
Ma after CAI, suggesting that they have indeed recorded common catastrophic event(s). In
fact, the closure ages estimated here probably result of the integration of a suite of events
occurring during the accretion and where primitive and differentiated bodies with variable
volatile contents contributed to the building of the proto-Earth.
The fraction of the proto-atmosphere removed during a giant impact is debated [45,70].
New modelling results based on the Moon-forming scenario with a fast-spinning Earth [71]
suggest that a large part (if not the whole) of the atmosphere could have been removed during
this major event [72]. Our reconstructed Xe budget indicates that about 23 % of the total
129Xe(I) produced by terrestrial 129I was left in the atmosphere after completion of the Earth.
The possibility that ≈ 77 % of initial volatile elements would have been lost before
atmosphere's closure is not unreasonable in regard to possible atmospheric erosion during
terrestrial accretion [45]. This mass balance implies that the proto-Earth could have hosted a
factor of ≈ 4 more volatiles than at present, if these volatiles were supplied before the last
giant impact event. Since the present-day terrestrial inventory of volatile element is estimated
to be equivalent to about 2±1 % of carbonaceous chondrite (CC) material [73], this mass
balance suggests that the proto-Earth could have been contributed by up to ≈ 10 % CC
material before the last giant impact. Alternatively, terrestrial volatiles could have been
supplied from bodies having Xe/I ratios higher than those observed presently in CC (e.g.,
cometary material?).
The conclusions drawn here have to be taken carefully as there all still too many shady
areas in the early accretional history of the Earth. Our model does not pretend to describe
such accretional processes, which will require a continuous accretion/degassing simulation.
The main result of this study is nevertheless that the age of the terrestrial atmosphere appears
closer to ≈ 40 Ma rather than to 100 Ma as previously thought.
5. Conclusion
The I-Pu-Xe system gives useful, yet not fully understood, time constraints "on
establishment of global chemical inventories" [17] for the Earth and, presumably, for the
Moon. After geological loss of Xe from the atmosphere, the closure ages derived from the I-
Xe and I-Pu-Xe systematics are more ancient than previously proposed and suggest major
13
forming events around 40 Ma (range 30-60 Ma) after CAI. A more comprehensive approach
of this chronology will need to integrate the I-Pu-Xe system into n-body simulations of
terrestrial accretion, parameterized with gain and loss of volatiles, especially during the giant
impact epoch.
Acknowledgments.
David Stevenson and Alex Halliday, organisers of the "Origin of the Moon" meeting,
are gratefully acknowledged as well as other participants of the meeting for fruitful
discussions. Remarks from Maïa Kuga and Eric Hébrard helped to build this model. We are
grateful to Allessandro Morbidelli, Seth Jacobson, and Patrick Michel for sharing exciting
ideas on the accretion of terrestrial planets, and to Jamie Gilmour and Sujoy Mukhopadhyay
for constructive reviews. This work was supported by the European Research Council under
the European Community's Seventh Framework Program (FP7/2007-2013 grant agreement
no. 267255 to B.M.). This is CRPG contribution #2308
14
References
1
2
3
Patterson, C. 1956 Age of meteorites and the earth. Geochim. Cosmochim. Acta 10,
230–237.
Amelin, Y., Krot, A. N., Hutcheon, I. D. & Ulyanov, A. A. 2002 Lead isotopic ages of
inclusions. Science 297, 1678–83.
chondrules
(doi:10.1126/science.1073950)
calcium-aluminum-rich
and
Bouvier, A. & Wadhwa, M. 2010 The age of the Solar System redefined by the oldest
Pb–Pb age of a meteoritic inclusion. Nat. Geosci. 3, 637–641. (doi:10.1038/ngeo941)
4 Mukhopadhyay, S. 2012 Early differentiation and volatile accretion recorded in deep-
mantle neon and xenon. Nature 486, 101–104.
5
6
7
8
9
10
Tucker, J. M., Mukhopadhyay, S. & Schilling, J.-G. 2012 The heavy noble gas
composition of the depleted MORB mantle (DMM) and its implications for the
preservation of heterogeneities in the mantle. Earth Planet. Sci. Lett. 355–356, 244–
254. (doi:http://dx.doi.org/10.1016/j.epsl.2012.08.025)
Parai, R., Mukhopadhyay, S. & Standish, J. J. 2012 Heterogeneous upper mantle Ne,
Ar and Xe isotopic compositions and a possible Dupal noble gas signature recorded in
basalts from the Southwest Indian Ridge. Earth Planet. Sci. Lett. 359-360, 227–239.
(doi:10.1016/j.epsl.2012.10.017)
Pető, M. K., Mukhopadhyay, S. & Kelley, K. a. 2013 Heterogeneities from the first
100 million years recorded in deep mantle noble gases from the Northern Lau Back-arc
Basin. Earth Planet. Sci. Lett. 369-370, 13–23. (doi:10.1016/j.epsl.2013.02.012)
Katcoff, S., Schaeffer, O. A. & Hastings, J. M. 1951 Half-Life of Iodine-129 and the
Age of the Elements. Phys. Rev. 82, 688–690.
Reynolds, J. H. 1960 Determination of the age of the elements. Phys. Rev. Lett. 4.
Porcelli, D. & Ballentine, C. J. 2002 Models for Distribution of Terrestrial Noble
Gases and Evolution of the Atmosphere. Rev. Mineral. Geochemistry 47, 411–480.
(doi:10.2138/rmg.2002.47.11)
11
Staudacher, T. & Allègre, C. J. 1982 Terrestrial xenology. Earth Planet. Sci. Lett. 60,
389–406. (doi:http://dx.doi.org/10.1016/0012-821X(82)90075-9)
12 Caffee, M. W., Hudson, G. B., Velsko, C., Huss, G. R., Alexander, E. C. & Chivas, A.
R. 1999 Primordial Noble Gases from Earth’s Mantle: Identification of a Primitive
Volatile Component. Science 285, 2115–2118. (doi:10.1126/science.285.5436.2115)
13 Marty, B. 1989 Neon and xenon isotopes in MORB: implications for the earth-
45–56.
atmosphere
(doi:http://dx.doi.org/10.1016/0012-821X(89)90082-4)
evolution.
Lett.
Earth
Planet.
Sci.
94,
15
14
Padhi, C. M., Korenaga, J. & Ozima, M. 2012 Thermal evolution of Earth with xenon
degassing: A self-consistent approach. Earth Planet. Sci. Lett. 341–344, 1–9.
(doi:http://dx.doi.org/10.1016/j.epsl.2012.06.013)
15 Wetherill, G. W. 1975 Radiometric Chronology of the Early Solar System. Annu. Rev.
Nucl. Sci. 25, 283–328. (doi:10.1146/annurev.ns.25.120175.001435)
16 Hohenberg, C. M., Podosek, F. A. & Reynolds, J. H. 1967 Xenon-Iodine Dating: Sharp
233–236.
Chondrites.
Science
156
,
Isochronism
(doi:10.1126/science.156.3772.233)
in
17
Podosek, F. A. & Ozima, M. 2000 The Xenon Age of the Earth. In Origin of the Earth
and Moon, pp. 63–72.
18 Tolstikhin, I., Marty, B., Porcelli, D. & Hofmann, A. 2013 Evolution of volatile species
in the earth’s mantle: a view from xenology. Geochim. Cosmochim. Acta In Press.
(doi:10.1016/j.gca.2013.08.034)
19 Allègre, C. J., Staudacher, T., Sarda, P. & Kurz, M. D. 1983 Constraints on evolution
of Earth’s mantle from rare gas systematics. Nature 303, 762–766.
20 Alexander, E. C., Lewis, R. S., Reynolds, J. H. & Michel, M. C. 1971 Plutonium-244:
837–840.
an Extinct Radioactivity.
Science
172
,
Confirmation
(doi:10.1126/science.172.3985.837)
as
21 Hudson, G. B., Kennedy, B. M., Podosek, F. A. & Hohenberg, C. M. 1989 The Early
Solar System Abundance of 244Pu as inferred from the St. Severin Chondrite. Proc.
19th Lunar Planet. Sci. Conf. , 547–557.
22 Kunz, J., Staudacher, T. & Allègre, C. J. 1998 Plutonium-Fission Xenon Found in
Earth’s Mantle. Science 280, 877–880. (doi:10.1126/science.280.5365.877)
23
Pepin, R. O. 1991 On the origin and early evolution of terrestrial planet atmospheres
and meteoritic volatiles. Icarus 92, 2–79. (doi:http://dx.doi.org/10.1016/0019-
1035(91)90036-S)
24
Pepin, R. O. 2000 On the isotopic composition of primordial xenon in terrestrial planet
atmospheres. Space Sci. Rev. 92, 371–395.
25 Takaoka, N. 1972 An interpretation of general anomalies of xenon and the isotopic
composition of primitive xenon. Mass Spectrosc. 20, 287–302.
26
Pepin, R. O. & Phinney, D. 1978 Components of xenon in the solar system
(Unpublished preprint).
27 Huss, G. R. & Lewis, R. S. 1995 Presolar diamond, SiC, and graphite in primitive
chondrites: Abundances as a function of meteorite class and petrologic type. Geochim.
Cosmochim. Acta 59, 115–160. (doi:10.1016/0016-7037(94)00376-W)
16
28 Gilmour, J. D. 2010 “Planetary” noble gas components and the nucleosynthetic history
380–393.
system material. Geochim. Cosmochim. Acta
of
(doi:http://dx.doi.org/10.1016/j.gca.2009.09.015)
solar
74,
29 Crowther, S. A. & Gilmour, J. D. 2013 The Genesis Solar Xenon Composition and its
Relationship to Planetary Xenon Signatures. Geochim. Cosmochim. Acta 123, 17–34.
(doi:10.1016/j.gca.2013.09.007)
30
Igarashi, G. 1995 Primitive xenon in the earth. AIP Conf. Proc. 341, 70–80.
(doi:10.1063/1.48751)
31 Ozima, M. & Podosek, F. A. 2002 Noble Gas Geochemistry: 2nd edition. Cambridge
University Press.
32
Pepin, R. O. & Porcelli, D. 2006 Xenon isotope systematics, giant impacts, and mantle
the early Earth. Earth Planet. Sci. Lett. 250, 470–485.
degassing on
(doi:http://dx.doi.org/10.1016/j.epsl.2006.08.014)
33 Tolstikhin, I. N. & Marty, B. 1998 The evolution of terrestrial volatiles: a view from
helium, neon, argon and nitrogen isotope modelling. Chem. Geol. 147, 27–52.
(doi:http://dx.doi.org/10.1016/S0009-2541(97)00170-8)
34 Yokochi, R. & Marty, B. 2005 Geochemical constraints on mantle dynamics in the
17–30.
Sci.
Hadean.
(doi:http://dx.doi.org/10.1016/j.epsl.2005.07.020)
Planet.
Earth
Lett.
238,
35
36
37
38
Srinivasan, B. 1976 Barites: anomalous xenon from spallation and neutron-induced
reactions. Earth Planet. Sci. Lett. 31, 129–141. (doi:http://dx.doi.org/10.1016/0012-
821X(76)90104-7)
Pujol, M., Marty, B., Burnard, P. & Philippot, P. 2009 Xenon in Archean barite: Weak
decay of 130Ba, mass-dependent isotopic fractionation and implication for barite
formation.
6834–6846.
(doi:http://dx.doi.org/10.1016/j.gca.2009.08.002)
Cosmochim.
Geochim.
Acta
73,
Pujol, M., Marty, B., Burgess, R., Turner, G. & Philippot, P. 2013 Argon isotopic
composition of Archaean atmosphere probes early Earth geodynamics. Nature 498, 87–
90.
Pujol, M., Marty, B. & Burgess, R. 2011 Chondritic-like xenon trapped in Archean
rocks: A possible signature of the ancient atmosphere. Earth Planet. Sci. Lett. 308,
298–306. (doi:http://dx.doi.org/10.1016/j.epsl.2011.05.053)
39 Holland, G., Lollar, B. S., Li, L., Lacrampe-Couloume, G., Slater, G. F. & Ballentine,
C. J. 2013 Deep fracture fluids isolated in the crust since the Precambrian era. Nature
497, 357–60. (doi:10.1038/nature12127)
40 Marrocchi, Y., Marty, B., Reinhardt, P. & Robert, F. 2011 Adsorption of xenon ions
onto defects in organic surfaces: Implications for the origin and the nature of organics
17
primitive meteorites. Geochim. Cosmochim. Acta
in
(doi:http://dx.doi.org/10.1016/j.gca.2011.07.048)
75,
6255–6266.
41 Ribas, I., Guinan, E. F., Güdel, M. & Audard, M. 2005 Evolution of the Solar Activity
over Time and Effects on Planetary Atmospheres. I. High-Energy Irradiances (1-1700
Å). Astrophys. J. 622, 680–694.
42 Hébrard, E. & Marty, B. 2014 Coupled noble gas–hydrocarbon evolution of the early
Earth atmosphere upon solar UV irradiation. Earth Planet. Sci. Lett. 385, 40–48.
(doi:10.1016/j.epsl.2013.10.022)
43 Zahnle, K., Arndt, N., Cockell, C., Halliday, A., Nisbet, E., Selsis, F. & Sleep, N. H.
2007 Emergence of a Habitable Planet. Space Sci. Rev. 129, 35–78.
(doi:10.1007/s11214-007-9225-z)
44 Chen, G. Q. & Ahrens, T. J. 1997 Erosion of terrestrial planet atmosphere by surface
Inter. 100, 21–26.
impact. Phys. Earth Planet.
motion after a
(doi:10.1016/S0031-9201(96)03228-1)
large
45 Genda, H. & Abe, Y. 2005 Enhanced atmospheric loss on protoplanets at the giant
impact phase in the presence of oceans. Nature 433, 842–4. (doi:10.1038/nature03360)
46 Canup, R. M. & Agnor, C. B. 2000 Accretion of the Terrestrial Planets and the Earth-
Moon System. In Origin of the Earth and Moon, pp. 113–130.
47 Blichert-Toft, J. & Albarède, F. 1994 Short-lived chemical heterogeneities in the
archean mantle with implications for mantle convection. Science 263, 1593–1596.
(doi:10.1126/science.263.5153.1593)
48 Coltice, N. 2005 The role of convective mixing in degassing the Earth’s mantle. Earth
Planet. Sci. Lett. 234, 15–25. (doi:http://dx.doi.org/10.1016/j.epsl.2005.02.041)
49 Gonnermann, H. M. & Mukhopadhyay, S. 2009 Preserving noble gases in a convecting
mantle. Nature 459, 560–3. (doi:10.1038/nature08018)
50 Korenaga, J. 2003 Energetics of mantle convection and the fate of fossil heat. Geophys.
Res. Lett. 30, 1437. (doi:10.1029/2003GL016982)
51
Frick, U., Mack, R. & Chang, S. 1979 Noble gas trapping and fractionation during
synthesis of carbonaceous matter. Proc. Lunar Sci. Conf. 10. , 1961–1972.
52 Bernatowicz, T. J. & Fahey, A. J. 1986 Xe isotopic fractionation in a cathodeless glow
discharge. Geochim. Cosmochim. Acta 50, 445–452.
53 Bernatowicz, T. J. & Hagee, B. E. 1987 Isotopic fractionation of Kr and Xe implanted
in solids at very low energies. Geochim. Cosmochim. Acta 51, 1599–1611.
54
Ponganis, K., Graf, T. & Marti, K. 1997 Isotopic fractionation in low-energy ion
implantation. J. Geophys. Res. 102, 19335–19343.
18
55 Hohenberg, C. M., Thonnard, N. & Meshik, A. 2002 Active capture and anomalous
adsorption: New mechanisms for the incorporation of heavy noble gases. Meteorit.
Planet. Sci. 37, 257–267. (doi:10.1111/j.1945-5100.2002.tb01108.x)
56 Déruelle, B., Dreibus, G. & Jambon, A. 1992 Iodine abundances in oceanic basalts:
for Earth dynamics. Earth Planet. Sci. Lett. 108, 217–227.
implications
(doi:10.1016/0012-821X(92)90024-P)
57 Armytage, R. M. G., Jephcoat, A. P., Bouhifd, M. A. & Porcelli, D. 2013 Metal–
silicate partitioning of iodine at high pressures and temperatures: Implications for the
Earth’s core and 129*Xe budgets. Earth Planet. Sci. Lett. 373, 140–149.
(doi:10.1016/j.epsl.2013.04.031)
58 Wänke, H., Dreibus, G. & Jagoutz, E. 1984 Mantle Chemistry and Accretion History of
the Earth. In Archaean Geochemistry (eds A. Kröner G. N. Hanson & A. M. Goodwin),
pp. 1–24. Springer.(doi:10.1007/978-3-642-70001-9_1)
59 Lugmair, G. W. & Marti, K. 1977 Sm-Nd-Pu timepieces in the Angra dos Reis
meteorite. Earth Planet. Sci. Lett. 35, 272–284.
60 Hudson, G. B., Kennedy, B. M., Podosek, F. A. & Hohenberg, C. M. 1989 The early
solar system abundance of 244Pu as inferred from the St. Severin chondrite. Proc. 19th
Lunar Planet. Sci. Conf. , 547–555.
61 Honda, M., Nutman, A. P. & Bennett, V. C. 2003 Xenon compositions of magmatic
zircons in 3.64 and 3.81 Ga meta-granitoids from Greenland – a search for extinct
rocks. Earth Planet. Sci. Lett. 207, 69–82.
244Pu
(doi:10.1016/S0012-821X(02)01147-0)
in ancient
terrestrial
62 Turner, G., Harrison, T. M., Holland, G., Mojzsis, S. J. & Gilmour, J. 2004 Extinct
244Pu in ancient zircons. Science 306, 89–91. (doi:10.1126/science.1101014)
63 Turner, G., Busfield, A., Crowther, S. A., Harrison, M., Mojzsis, S. J. & Gilmour, J.
2007 Pu–Xe, U–Xe, U–Pb chronology and isotope systematics of ancient zircons from
Earth
Western
491–499.
(doi:10.1016/j.epsl.2007.07.014)
Australia.
Lett.
Planet.
Sci.
261,
64
Porcelli, D. & Pepin, R. O. 2000 Rare Gas Constraints on Early Earth History. In
Origin of the Earth and Moon, pp. 435–458.
65 Yin, Q., Jacobsen, S. B., Yamashita, K., Blichert-Toft, J., Télouk, P. & Albarède, F.
2002 A short timescale for terrestrial planet formation from Hf-W chronometry of
meteorites. Nature 418, 949–52. (doi:10.1038/nature00995)
66 Righter, K. & Shearer, C. K. 2003 Magmatic fractionation of Hf and W: constraints on
the timing of core formation and differentiation in the Moon and Mars. Geochim.
Cosmochim. Acta 67, 2497–2507. (doi:10.1016/S0016-7037(02)01349-2)
67 Halliday, A. N. 2004 Mixing, volatile loss and compositional change during impact-
driven accretion of the Earth. Nature 427, 505–9. (doi:10.1038/nature02275)
19
68 Kleine, T., Palme, H., Mezger, K. & Halliday, A. N. 2005 Hf-W chronometry of lunar
metals and the age and early differentiation of the Moon. Science 310, 1671–4.
(doi:10.1126/science.1118842)
69 Touboul, M., Kleine, T., Bourdon, B., Palme, H. & Wieler, R. 2007 Late formation and
prolonged differentiation of the Moon inferred from W isotopes in lunar metals. Nature
450, 1206–1209.
70 Genda, H. & Abe, Y. 2003 Survival of a proto-atmosphere through the stage of giant
(doi:10.1016/S0019-
the mechanical aspects.
Icarus 164, 149–162.
impacts:
1035(03)00101-5)
71 Ćuk, M. & Stewart, S. T. 2012 Making the Moon from a fast-spinning Earth: a giant
1047–52.
despinning.
resonant
Science
338,
impact
(doi:10.1126/science.1225542)
followed
by
72 Lock, S. J. & Stewart, S. t. 2013 Atmospheric loss during high angular momentum
giant impacts (44th LPSC). , 2608.
73 Marty, B. 2012 The origins and concentrations of water, carbon, nitrogen and noble
56–66.
gases
(doi:http://dx.doi.org/10.1016/j.epsl.2011.10.040)
313–314,
Planet.
Earth.
Earth
Lett.
on
Sci.
Porcelli, D. & Turekian, K. K. 2003 The History of Planetary Degassing as Recorded
by Noble Gases. In Treatise on Geochemistry, Volume 4 (eds H. D. Holland & K. K.
Turekian), pp. 281–318. Elsevier.
20
74
Tables
Table 1: Input parameters of the model
parameter
IINI
(129I/127I)INI
238UINI
(244Pu/238U)INI
244PuINI
130XeINI
(124Xe/130Xe)INI
(126Xe/130Xe)INI
(128Xe/130Xe)INI
(129Xe/130Xe)INI
(131Xe/130Xe)INI
(132Xe/130Xe)INI
(134Xe/130Xe)INI
(136Xe/130Xe)INI
literature note
3-13
a
ppb
mol/mol 1.1x10-4 b
ppb
c
mol/mol 6.8x10-3 d
e
mol
33-41
n.d.
f
mol
mol/mol 0.02928 g
-
-
-
-
-
-
-
-
-
-
-
-
-
-
0.02534
0.5083
6.286
4.996
6.047
2.126
1.657
degassing parameters
φ1
φ2
φ3
escape parameters
b
c
d
Ma-1
-
-
Ma-1
-
-
n.d
n.d
n.d
2.53
0.85
n.d
instantaneous
fractionation
%.amu-1 1.0-1.5
h
-
-
i
-
j
notes:
adjusted value*
6.4
1.1x10-4
40
6.8x10-3
3.2x1015
1.44x1013
0.02928
0.02534
0.5083
6.286
4.996
6.047
2.126
1.657
0.065
0.0011
0.000073
2.53
0.85
1.2270
1.3
* bold values are adjusted parameters using multiple runs of the model.
a, b: See section 3.(a)
21
c, d, e: See section 3.(b)
g: See section 3.(c)
h, i, j: See section 3.(d)
22
Table 2: Reference solution of the model compared to values from the literature
parameter
XeATM
(124Xe/130Xe)ATM
(126Xe/130Xe)ATM
(128Xe/130Xe)ATM
(129Xe/130Xe)ATM
(131Xe/130Xe)ATM
(132Xe/130Xe)ATM
(134Xe/130Xe)ATM
(136Xe/130Xe)ATM
dimension
mol
mol/mol
-
-
-
-
-
-
-
literature
1.537x1013
0.02337
0.0218
0.4715
6.496
5.213
6.607
2.563
2.176
129Xe(I)ATM,CORR
(129Xe(I)/136Xe(Pu))ATM
(129Xe(I)/136Xe(Pu))ATM,CORR -
mol
mol/mol
6.5-7.1
-
XeMANT
(124Xe/130Xe)MANT
(126Xe/130Xe)MANT
(128Xe/130Xe)MANT
(129Xe/130Xe)MANT
(131Xe/130Xe)MANT
(132Xe/130Xe)MANT
(134Xe/130Xe)MANT
(136Xe/130Xe)MANT
notes:
mol
mol/mol
-
-
-
-
-
-
-
reference solution
1.5184x1013
0.0232
0.0217
0.470
6.522
5.21
6.60
2.55
2.17
3.71x1012
6.99
21.75
2.0356x1012
0.0293
0.0253
0.508
9.00
5.18
6.73
2.86
2.50
note
a
-
-
-
-
-
-
-
-
b
c
d
*Bold values are ajusted with multiple runs of the model (see Section 3.4).
a- The total inventory and isotopic ratios of Xe are from [31].
b- Amount of 129Xe in the atmosphere coming from the decay of 129I and corrected for loss.
c- Ratio of radioactivity products uncorrected for loss from [74].
d- Ratio corrected for loss.
23
Figure captions
Fig. 1: Relationship between the evolution of the solar EUV flux with time and the
progressive isotopic fractionation of atmospheric xenon. (a) Evolution of the Solar flux with
time. Figure and data are from [41]. The wavelength of ionization of Xe atoms corresponds to
the range 920-1200 Å. (b) Progressive isotopic fractionation of atmospheric Xe from
cosmochemical components [10] to modern atmosphere [31]. Ancient rocks record
intermediate isotopic compositions: 1 Ga-old barites [35,36], Quartz samples [37,38],
Proterozoic deep fracture fluids [39].
Fig. 2: (a) Schematic explanation of the model and evolution of the budget of
atmospheric Xe over time. The model is built with three boxes: solid Earth, atmosphere and
space. Some of Xe isotopes are produced by radioactive decay of 129I, 244Pu and 238U (see
text). After a time interval Δt (40 Ma, one outcome of the model) the Earth begins to retain its
volatile elements and accumulates xenon degassed from the Solid Earth to the atmosphere
without isotopic fractionation. Xenon atoms escape from the atmosphere to the outer space
with isotopic fractionation. The evolution of the budget of atmospheric xenon shows the
progressive escape of Xe atoms with time. The escape lasts until the end of the Archean eon
(t=2 Ga). At this time, the abundance has almost reached the current abundance of xenon in
the atmosphere. Further degassing permits to complete the amount of xenon atoms. (b)
Isotopic spectrum of xenon relative to the current isotopic composition of the Earth's
atmosphere using 130Xe as a reference isotope. Xe-U is the starting isotopic composition
(circles). The fractionated Archean atmosphere (around 1%.amu-1) is shown with squares and
the "artificial" current isotopic composition of the reference solution is shown with stars. The
current isotopic composition is reproduced within 0.7% or better.
Fig. 3: (a) Evolution of the atmospheric content of 129Xe derived from the decay of 129I
with time. The non-corrected amount gave a closure age of 98 Ma for the I-Xe system. After
correction for subsequent loss, the age becomes 41 Ma. (b) Evolution with time of the ratio of
radioactive products in the atmosphere (129Xe(I) and 136Xe(Pu)). The non-corrected ratio gave
a closure age of 66 Ma for the I-Pu-Xe system. After correction, the age becomes 34 Ma in
agreement with the time of closure of the I-Xe system and with the closure age of the mantle
given by the mantle samples [4,32].
24
1-20 Å
20-100 Å
100-360 Å
360-920 Å
920-1200 Å
0.1
1
t (Gyr)
8
Cosmochemical components
Archean barites
Quartz
Deep fracture fluids
Modern atmosphere
(a)
1000
)
1
=
n
u
S
(
x
u
l
f
e
v
i
t
a
l
e
R
100
10
1
0.1
(b)
)
‰
(
)
e
X
0
3
1
/
e
X
i
(
δ
40
30
20
10
0
0.1
1
t (Gyr)
Figure 1
10
129Xe(I) (mol.)
(a)
6x1012
I
I
I
I
I
N
I
N
N
I
I
=
3
p
I
=
6
=
1
.
3
4
p
p
b
p
p
p
b
b
Escape-corrected values
5x1012
4x1012
3x1012
2x1012
1x1012
0
20
60
80
41 Ma
129Xe(I)/136Xe(Pu)
(b)
100
98 Ma
Present-day
120
t (Ma)
60
50
40
30
20
10
I
I
N
I=
3
p
p
b
I
I
N
I
=
1
I
I
N
I
=
6
.
4
p
p
b
3
p
p
b
Escape-corrected values
0
20
40
34 Ma
60
66 Ma
80
Figure 3
Present-day
100
120
t (Ma)
|
1907.13425 | 1 | 1907 | 2019-07-31T11:35:33 | He I $\lambda$ 10830 \r{A} in the transmission spectrum of HD 209458 b | [
"astro-ph.EP"
] | Context: Recently, the He I triplet at 10830 \r{A} has been rediscovered as an excellent probe of the extended and possibly evaporating atmospheres of close-in transiting planets. This has already resulted in detections of this triplet in the atmospheres of a handful of planets, both from space and from the ground. However, while a strong signal is expected for the hot Jupiter HD 209458 b, only upper limits have been obtained so far. Aims: Our goal is to measure the helium excess absorption from HD 209458 b and assess the extended atmosphere of the planet and possible evaporation. Methods: We obtained new high-resolution spectral transit time-series of HD 209458 b using CARMENES at the 3.5 m Calar Alto telescope, targeting the He I triplet at 10830 \r{A} at a spectral resolving power of 80 400. The observed spectra were corrected for stellar absorption lines using out of transit data, for telluric absorption using the molecfit software, and for the sky emission lines using simultaneous sky measurements through a second fibre. Results: We detect He I absorption at a level of 0.91 $\pm$ 0.10 % (9 $\sigma$) at mid-transit. The absorption follows the radial velocity change of the planet during transit, unambiguously identifying the planet as the source of the absorption. The core of the absorption exhibits a net blueshift of 1.8 $\pm$ 1.3 km s$^{-1}$. Possible low-level excess absorption is seen further blueward from the main absorption near the centre of the transit, which could be caused by an extended tail. However, this needs to be confirmed. Conclusions: Our results further support a close relationship between the strength of planetary absorption in the helium triplet lines and the level of ionising, stellar X-ray and extreme-UV irradiation. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. Alonso-Floriano_2019
August 1, 2019
c(cid:13)ESO 2019
He i λ 10830 Å in the transmission spectrum of HD 209458 b
F. J. Alonso-Floriano1, I. A. G. Snellen1, S. Czesla2, F. F. Bauer3, M. Salz2, M. Lampón3, L. M. Lara3, E. Nagel2,
M. López-Puertas3, L. Nortmann4, 5, A. Sánchez-López3, J. Sanz-Forcada6, J. A. Caballero6, A. Reiners7, I. Ribas8, 9,
A. Quirrenbach10, P. J. Amado2, J. Aceituno11, G. Anglada-Escudé12, V. J. S. Béjar5, 6, M. Brinkmöller10,
A. P. Hatzes13, Th. Henning14, A. Kaminski10, M. Kürster14, F. Labarga15, D. Montes15, E. Pallé5, 6,
J. H. M. M. Schmitt3, and M. R. Zapatero Osorio16
9
1
0
2
l
u
J
1
3
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
2
4
3
1
.
7
0
9
1
:
v
i
X
r
a
1 Leiden Observatory, Leiden University, Postbus 9513, 2300 RA, Leiden, The Netherlands
2 Hamburger Sternwarte, Universität Hamburg, Gojenbergsweg 112, 21029 Hamburg, Germany
3 Instituto de Astrofísica de Andalucía (IAA-CSIC), Glorieta de la Astronomía s/n, 18008 Granada, Spain
4 Instituto de Astrofísica de Canarias (IAC), Calle Vía Lactea s/n, E-38200 La Laguna, Tenerife, Spain
5 Departamento de Astrofísica, Universidad de La Laguna, 38026 La Laguna, Tenerife, Spain
6 Centro de Astrobiología (CSIC-INTA), ESAC, Camino bajo del castillo s/n, 28692 Villanueva de la Cañada, Madrid, Spain
7 Institut für Astrophysik, Georg-August-Universität, 37077 Göttingen, Germany
8 Institut de Ciències de l'Espai (CSIC-IEEC), Campus UAB, c/ de Can Magrans s/n, 08193 Bellaterra, Barcelona, Spain
9 Institut d'Estudis Espacials de Catalunya (IEEC), 08034 Barcelona, Spain
10 Landessternwarte, Zentrum für Astronomie der Universität Heidelberg, Königstuhl 12, 69117 Heidelberg, Germany
11 Centro Astronónomico Hispano Alemán, Observatorio de Calar Alto, Sierra de los Filabres, E-04550 Gérgal, Spain
12 School of Physics and Astronomy, Queen Mary, University of London, 327 Mile End Road, London, E1 4NS, UK
13 Thüringer Landessternwarte Tautenburg, Sternwarte 5, 07778 Tautenburg, Germany
14 Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidelberg, Germany
15 Departamento de Física de la Tierra y Astrofísica and IPARCOS-UCM (Intituto de Física de Partículas y del Cosmos de la UCM),
Facultad de Ciencias Físicas, Universidad Complutense de Madrid, E-28040, Madrid, Spain
16 Centro de Astrobiología (CSIC-INTA), Carretera de Ajalvir km 4, E-28850 Torrejón de Ardoz, Madrid, Spain
Received 29 May 2019; Accepted 30 July 2019
ABSTRACT
Context. Recently, the He i triplet at 10830 Å has been rediscovered as an excellent probe of the extended and possibly evaporating
atmospheres of close-in transiting planets. This has already resulted in detections of this triplet in the atmospheres of a handful of
planets, both from space and from the ground. However, while a strong signal is expected for the hot Jupiter HD 209458 b, only upper
limits have been obtained so far.
Aims. Our goal is to measure the helium excess absorption from HD 209458 b and assess the extended atmosphere of the planet and
possible evaporation.
Methods. We obtained new high-resolution spectral transit time-series of HD 209458 b using CARMENES at the 3.5 m Calar Alto
telescope, targeting the He i triplet at 10830 Å at a spectral resolving power of 80 400. The observed spectra were corrected for stellar
absorption lines using out of transit data, for telluric absorption using the molecfit software, and for the sky emission lines using
simultaneous sky measurements through a second fibre.
Results. We detect He i absorption at a level of 0.91± 0.10% (9 σ) at mid-transit. The absorption follows the radial velocity change
of the planet during transit, unambiguously identifying the planet as the source of the absorption. The core of the absorption exhibits
a net blueshift of 1.8± 1.3 km s−1. Possible low-level excess absorption is seen further blueward from the main absorption near the
centre of the transit, which could be caused by an extended tail. However, this needs to be confirmed.
Conclusions. Our results further support a close relation between the strength of planetary absorption in the helium triplet lines and
the level of ionising, stellar X-ray and extreme-UV irradiation.
Key words. planets and satellites: atmospheres -- planets and satellites: individual (HD 209458 b) -- techniques: spectroscopic --
Infrared: planetary systems
1. Introduction
Atmospheric mass loss can have a profound influence on the
evolution of a planet, including the early evolution of the Earth
(Lammer et al. 2008) and other Solar System planets. Hot
Jupiters, gas giants in tight orbits, are ideal objects to study at-
mospheric mass loss in detail. A large, extended atmosphere was
first detected around HD 209458 b targeting the Lyα line (Vidal-
Madjar et al. 2003), showing a deep transit signal pointing at at-
mospheric escape. While the extended atmosphere around this
planet was subsequently also detected in carbon and oxygen
(Vidal-Madjar et al. 2004), Lyα has remained a primary tool to
study planetary atmospheres and their escape (e.g., Lecavelier
des Etangs et al. 2010; Kulow et al. 2014; Ehrenreich et al. 2015;
Lavie et al. 2017; Bourrier et al. 2017, 2018). Most notably,
Ehrenreich et al. (2015) and later Lavie et al. (2017) showed that
the hot Neptune GJ 436 b is surrounded by a giant comet-like
cloud of hydrogen, extending far beyond the Roche radius.
Article number, page 1 of 7
A&A proofs: manuscript no. Alonso-Floriano_2019
Unfortunately, Lyα is located in the ultraviolet part of the
spectrum and currently only accessible using the Hubble Space
Telescope (HST). Furthermore, observations are strongly ham-
pered by interstellar absorption and air glow emission originat-
ing from a halo of hydrogen atoms around the Earth, contami-
nating the profile of the Lyα line. Almost two decades ago, Sea-
ger & Sasselov (2000) proposed the He i transition at 10830 Å
(in air), a triplet of absorption lines of a metastable state of he-
lium, as a good candidate to study the extended atmospheres of
close-in giant planets. Moutou et al. (2003) studied this helium
transition during a transit of HD 209458 b -- observed with the
spectroscopic mode of ISAAC on the VLT -- and obtained a 3 σ
upper limit of 0.5% for a 3 Å bandwidth.
Recently, three independent groups detected He i λ 10830 Å
in exoplanet atmospheres. Spake et al. (2018) detected the triplet
in WASP-107 b using spectrophotometry with HST/WFC3.
Spectroscopic detections at high-resolution were made in
WASP-69 b, HAT-P-11 b, and HD 189733 b (Nortmann et al.
2018; Allart et al. 2018; Salz et al. 2018, of 3.59± 0.19%,
∼1.2± 0.2%1, and 1.04± 0.09%, maximum reached absorp-
tions respectively). Another spectrophotometrical detection with
HST/WFC3 was later made in HAT-P-11 b (Mansfield et al.
2018) and ground spectroscopic observations of WASP-107 b
(Allart et al. 2019, 7.92± 1.00%) confirmed the earlier HS T de-
tection and revealed detailed information on the helium escape
as in the form of a cometary-like tail.
The ground-based measurements are particularly informa-
tive, since they are conducted using high-dispersion spec-
troscopy -- in all four cases with the CARMENES spectrograph
on the 3.5m Calar-Alto Telescope (Quirrenbach et al. 2016,
2018). At a resolving power of 80 400, details on the radial ve-
locity distribution of the helium gas are available. For example,
Nortmann et al. (2018) measured blueshifts of several kilometers
per second, which in combination with post-transit absorption,
was interpreted as the escape of part of the atmosphere trailing
behind the planet in comet-like form.
Nortmann et al. (2018) also provided He i upper limits for
three additional planets, HD 209458 b, Kelt-9 b, and GJ 436 b, of
0.84%, 0.33%, and 0.41%, respectively (90% confidence limits).
The wide range of He i absorption levels found for close-in plan-
ets is likely linked to the level of stellar X-ray and extreme-UV
(5 -- 504 Å) irradiation of the host star, populating the metastable
23S helium triple state (Nortmann et al. 2018). A similar relation
between the He triplet and ionizing XUV (λ < 504 Å) radiation
was previously observed in stellar coronae (e.g., Sanz-Forcada
& Dupree 2008).
The observations presented in Nortmann et al. (2018) for
HD 209458 b were performed under sub-optimal conditions, re-
sulting in poor data quality that hampered the search for helium
(see Sect. 4). In this work, we present new CARMENES transit
observations targeting the He i λ 10830 Å triplet, now resulting
in a firm detection at 0.91± 0.10%. In Section 2 we present the
observations, in Section 3 the data analysis, in Section 4 the dis-
cussion on the results, and we conclude in Section 5.
1 The error is an approximation obtained from the average transmis-
sion spectrum provided by Allart et al. (2018).
Article number, page 2 of 7
Table 1. Parameters of the exoplanet system HD 209458.
Parameter
α [J2000]
δ [J2000]
d
V
J
sys
K(cid:63)
R(cid:63)
M(cid:63)
Teff
Fa
XUV
Porb
T0 [HJD]
td
i
a
e cos ω
RP
MP
gP
Teq
Kb
P
Value
Referencec
22:03:10.77
+18:53:03.5
48.37 (12) pc
7.63 (1) mag
6.59 (2) mag
-- 14.7652 (16) km s−1
84.67 (70) m s−1
1.155+0.014
−0.016 R(cid:12)
1.119 (33) M(cid:12)
6065 (50) K
1.004 (284) W m−2
3.52474859 (38) d
2452826.628521 (87) d
183.89 (3.17) min
−0.00047 au
86.71 (5) deg
0.04707+0.00046
0.00004 (33)
1.359 +0.016
−0.019 RJ
0.685 +0.015
−0.014 MJ
9.18 (1) m s−2
1449 (12) K
144.9+5.4−5.3 km s−1
1
1
1
2
3
4
5
5
5
5
6
7
7
8
5
5
9
5
5
5
5
6
Notes. (a) X-ray and EUV (5 -- 504 Å) flux at the distance of the planet's
semi-major axis, derived from coronal models . (b) Derived from or-
bital parameters. (c) References. (1) Gaia Collaboration et al. (2018),
(2) Høg et al. (2000), (3) Skrutskie et al. (2006), (4) Mazeh et al.
(2000), (5) Torres et al. (2008), (6) This work, (7) Knutson et al. (2007),
(8) Wright et al. (2011), (9) Crossfield et al. (2012).
2. Observations
We observed the system HD 209458 with CARMENES on 5
September 20182. The planet is an archetypal hot Jupiter, and
the first known transiting system of its kind (Charbonneau et
al. 2000; Henry et al. 2000; Table 1). As a bright nearby sys-
tem, it has been the subject of many observational studies tar-
geting its transmission spectrum (e.g. Charbonneau et al. 2002;
Deming et al. 2005b; Snellen et al. 2008, 2010; Hoeijmakers et
al. 2015; Hawker et al. 2018), its hydrogen upper atmosphere
(Vidal-Madjar et al. 2003; Ben-Jaffel 2007; Ehrenreich et al.
2008), and its emission spectrum (Deming et al. 2005a; Knut-
son et al. 2008; Swain et al. 2009; Schwarz et al. 2015; Brogi et
al. 2017). This makes it one of the best studied exoplanets and a
benchmark for atmospheric evaporation modelling.
CARMENES consists of two spectrograph channels fed by
fibres connected to the front-end mounted on the telescope. One
of the CARMENES spectrographs, dubbed VIS channel, covers
the optical wavelength range ∆λ = 520 -- 960 nm over 55 orders.
The other, the NIR channel, covers the near-infrared wavelength
2 Program H18-3.5-022, P.I. S. Czesla. After one year from the obser-
vation date, the reduced spectra can be downloaded from the Calar Alto
archive, http://caha.sdc.cab.inta-csic.es/calto/
F. J. Alonso-Floriano et al.: He i in the transmission spectrum of HD 209458 b
Fig. 1. Change in airmass (top), precipitable water vapour (mid-
dle), and signal-to-noise ratio (bottom) during observations (mid-transit
∼ 00:39 UT). Vertical dashed lines indicate the first, second, third, and
fourth contact, respectively. The lack of data around −2.3 h was due to
a lost in the tracking system. The post-transit observations not included
in the analysis are shown in grey. The out-of-transit red diamonds were
used to compute the master spectrum of Fig. 2.
Fig. 2. The master spectrum in the vicinity of the He i triplet (blue
line). The vertical lines indicate the positions of the three He i lines.
The grey line indicates the average telluric absorption spectrum and the
sky emission lines removed from the data. The masked cores of telluric
absorption lines in Fig. 3 are indicated by thicker red lines. The latter
were highly variable during the night. The stellar line blueward from
the He triplet is Si i.
range ∆λ = 960 -- 1710 nm over 28 orders. The resolving power is
R = 94 600 in the VIS channel and R = 80 400 in the NIR chan-
nel. Two fibres are connected to each channel: fibre A is used
for the target, and fibre B for the sky. The observations were ob-
tained in service mode and consisted of 91 exposures of 198 s,
starting at 21:39 UT and ending at 03:47 UT on the night starting
on 5 September 2018, corresponding to a planet orbital phase
range of φ = -- 0.036 to +0.037. The tracking of the target was
lost during the pre-transit phase from 22:12 UT to 22:39 UT
(φ = -- 0.029 to -- 0.024, Fig. 1). We did not include these spectra
in our analysis. A typical continuum signal-to-noise ratio (S/N)
of ∼95 per spectrum was reached around the He i triplet. Since
the quality of the data decreased significantly during the post-
transit, down to a S/N of ∼40 -- 60, we also discarded the last eight
spectra (from 03:14 UT onward).
During the observing run, the airmass ranged between 1.05
and 2.11. We regularly updated the atmospheric dispersion cor-
rector, requiring a re-acquisition of the target, each time taking
about two minutes (c.f. Seifert et al. 2012). The update was done
every ∼ 40 min when the target was observed at altitudes >70◦
(airmass <1.06), every ∼ 30 min at altitudes between 70◦ and 50◦
(airmass 1.06 -- 1.30), and every ∼15 min at lower altitudes (air-
mass >1.30). During the course of the observation, the column of
precipitable water vapour (PWV) towards the target (measured
from the spectra) decreased from 10.3 to 3.8 mm.
3. Data analysis and results
The observed spectra were reduced using the CARMENES
pipeline Caracal v2.10 (Zechmeister et al. 2014; Caballero et
al. 2016). The pipeline provides a vacuum wavelength solution,
which we converted into air wavelengths, as used in the remain-
der of the paper.
The standard bad pixels mask of the NIR detector included
in Caracal v2.10 did not sufficiently correct for hot pixel effects.
In the data of fibre B, we included in the mask the left and right
neighbours around the hot pixels in the dispersion direction of
the detector, while in fibre A we also included the top and bot-
tom neighbours. A pair of remaining bad pixels present in the
extracted spectra were removed manually (pixels 482 and 489
corresponding to 10829.357 and 10829.757 Å). In one of the ob-
served spectra, taken at 00:13:58 UT, irregular variations in the
continuum were recorded for some orders. However, they did not
affect the spectral order containing the He i triplet.
Subsequently, a similar process as in Nortmann et al. (2018)
and Salz et al. (2018) was followed to remove the stellar and
telluric absorption and the sky emission lines from the spectra
around the He i triplet. We used version 1.5.9 of the molecfit
software (Smette et al. 2015; Kausch et al. 2015) to remove the
telluric absorption lines. However, as explained by Shulyak et
al. (2019), the telluric removal depends strongly on the S/N of
the data and atmospheric humidity, which might cause artefacts
in the core of the telluric lines after correction. Therefore, the
cores of the telluric lines could not be adequately corrected, and
these regions were masked out (see Figs. 2 and 3). We subtracted
the sky emission lines using the sky spectra from fibre B as ex-
plained in Salz et al. (2018). The fibre B spectra were extracted
in the same way than fibre A and corrected for cosmic rays by
fitting the temporal variation of each wavelength element with a
high-order polynomial, and substituting those values that deviate
by more than 5 σ.
We then normalised the stellar spectra by fitting their con-
tinuum, removing the cosmic-rays as in fibre B, and shifting the
spectra to the stellar rest-frame using the barycentric and sys-
temic velocities. A master spectrum was created (Fig. 2) com-
bining the out-of-transit spectra with weights at each wavelength
step based on the S/N following wi
λ , where xλ is
λ
the S/N at the wavelength, and i is the consecutive number of
the spectrum. Subsequently, each normalised spectrum was di-
vided by the master spectrum, resulting in an array of residuals
(wavelength vs. time) shown in Fig. 3.
λ /(cid:80)
= xi 2
i xi 2
Article number, page 3 of 7
−0.04−0.020.000.020.04406080100120−0.04−0.020.000.020.04Orbital phase406080100120S/N24681012PWV [mm]1.01.21.41.61.82.02.2Airmass−3−2−10123Time since mid−transit [h]10820108251083010835108400.50.60.70.80.91.01.11082010825108301083510840Wavelength (Å)0.50.60.70.80.91.01.1Flux Norm.A&A proofs: manuscript no. Alonso-Floriano_2019
Fig. 3. Time sequence of the residual spectra around the He i triplet,
after removal of the stellar and telluric absorption and the sky emission
lines, with the relative flux indicated in grey scale. The radial velocity in
the stellar rest-frame is on the horizontal axis, and orbital phase on the
vertical axis. The corresponding wavelengths in the rest-frame of the
star are indicated in the top axis. The area around orbital phase −0.03
is blanked out due to the lack of observations. The horizontal dashed
lines indicate the first, second, third and fourth contact of the transit,
respectively. A helium signal is visible (darker grey scales), following
the planet velocity (slanted lines) from −16 to +16 km s−1.
3.1. Transmission spectrum and light curve
Figure 3 shows the time sequence of the residual spectra around
the He i triplet after removal of the stellar and telluric absorp-
tion and the sky emission lines. A helium signal is visible during
transit following the planet radial velocity. This provides strong
evidence that the helium absorption has a planetary origin. In the
spectra obtained during mid-transit, a tentative extra absorption
feature is visible directly blueward from the main helium absorp-
tion, which we discuss below. The vertical features redward of
He i (60 and 120 km s−1) are masked areas corresponding to the
centres of strong, highly variable telluric lines. Residuals blue-
ward (−80 km s−1) from He i are from stellar Si i (Fig. 2) and pos-
sibly caused by the Rossiter-McLaughlin effect (see Sect. 4.1).
We shifted the residual spectra into the planet rest-frame and
computed the average transmission spectrum of HD 209458 b
using the 33 spectra collected between the second and third
contact of the transit (φ = -- 0.013 to +0.013, Fig. 4). The peak
value of the average absorption signal in the core of the two
strongest and blended lines of the He i triplet was at the level of
0.91 ± 0.10 %. The average absorption level over a bandwidth of
0.30 Å centred at the absorption peak was 0.71 ± 0.06 %. In both
cases, the absorption values and uncertainties were determined
using the bootstrap method (Fig. 5) as in Salz et al. (2018). The
third and weakest line of the triplet is not detected (see below).
The shape of the signal shows an asymmetry on the blue side,
which can also be seen at mid-transit in the two-dimensional ar-
ray of Fig. 3. We further discuss the transmission line profile in
Sect. 4.1.
We measure the peak of the helium absorption to be
blueshifted by 1.8± 1.3 km s−1 with respect to the restframe of
the planet. This is compatible with the blueshift of 2± 1 km s−1
observed for carbon monoxide by Snellen et al. (2010). How-
ever, this absorption is likely originated from a different layer in
the atmosphere of the planet.
Article number, page 4 of 7
Fig. 4. Average transmission spectrum around the He i triplet in the
planet rest-frame. The positions of the three helium lines are marked by
vertical lines. The red curve is the best-fit Parker wind model obtained
for a temperature of 6000 K and a mass-loss rate of 4.2×109 g s−1. The
blue line indicates a model for a tentative blueward component centred
at around −13 km s−1. The magenta dashed curve is the combination of
the red and blue lines.
Fig. 5. Histograms of the mean absorption levels in the average trans-
mission spectra generated in our bootstrap analysis when considering
only spectra between the second and third contact (left) and a control
sample including only the out-of-transit observations (right).
We constructed the light curve of the He i signal by mea-
suring the average absorption per spectrum between -- 13.3 and
+12.0 km s−1 (i.e., between 10829.814 and 10830.729 Å) in the
planet rest-frame, which is shown in Fig. 6. The average in-
transit absorption signal was ∼0.44%, about a factor two smaller
than the peak transmission signal as shown in Fig. 4 due to the
relatively wide integration band. There is no evidence for a pre-
or post-transit absorption signal.
3.2. Modelling of the helium triplet absorption
A one-dimensional isothermal hydrodynamic and spherically
symmetric model was used to calculate the He 23S density
in the planet upper atmosphere, similar to that developed by
−0.03−0.02−0.010.000.010.020.03Orbital phase −150−100−50050100150Radial velocity [km s−1]1082610828108301083210834Wavelength [Å]−0.015−0.010−0.0050.0000.0050.0100.01510827108281082910830108311083210833−1.0−0.8−0.6−0.4−0.20.00.20.410827108281082910830108311083210833Wavelength (Å)−1.0−0.8−0.6−0.4−0.20.00.20.4(cid:125)(cid:125)(cid:125)(cid:125)(cid:125)Fin/Fout−1 [%]−1.0−0.8−0.6−0.4−0.20.00.2Absorption level [%]050100150Histogram DensityF. J. Alonso-Floriano et al.: He i in the transmission spectrum of HD 209458 b
4. Discussion
Strong absorption of the upper atmosphere has previously been
detected in the transmission spectrum of HD 209458 b in Lyα
(Vidal-Madjar et al. 2003), atomic carbon and oxygen (Vidal-
Madjar et al. 2004), and magnesium (Vidal-Madjar et al. 2013),
at the 5 -- 10% level. The He i absorption presented in this work is
significantly lower, as expected due to the low fraction of helium
atoms in the excited state required to produce the absorption line
(Sect. 4.2). However, the mass-loss rates derived from our He i
analysis agree with those obtained from Lyα (see Sect. 3.2).
Previous observations by Moutou et al. (2003) and Nort-
mann et al. (2018) provided upper limits on the detection of He i
for HD 209458 b, which are consistent with the result presented
here. Moutou et al. derived an upper limit of 0.5% for a 3 Å band-
width. Assuming this bandwidth, our He i absorption detection
(0.91± 0.10%, FWHM∼0.4 Å) corresponds to a value of 0.12%.
In the case of Nortmann et al., the study of helium absorption
was hampered by the poor data quality. The two datasets used
in their analyses exhibit significantly lower signal-to-noise ratio
than our observations. Their data were obtained with the same
instrument, however, before an extensive intervention of the NIR
channel in November 2016, which improved the thermal man-
agement of the channel (Quirrenbach et al. 2018) and led to a
significant improvement in the achievable data quality. In addi-
tion, the observational settings were not optimal, viz., the atmo-
spheric dispersion corrector was not properly updated and cali-
bration images were obtained during transit resulting in a ∼30%
loss of signal for one of the nights. Nonetheless, the data of both
nights showed hints of absorption at the He i position (see panels
C and D on their Fig. S10), for which Nortmann et al. retrieved
an upper limit of 0.84% (90% confidence level).
If the helium signal were to be pursued with WFC3 on the
Hubble Space Telescope, its spectral resolution of 98 Å would
result in a transmission signal of ∼10 ppm. Even with NIRSPEC
(Dorner et al. 2016) on the soon to be launched James Webb
Space Telescope (JWST), this signal will only be at the ∼300 ppm
level over one resolution element, at the highest resolving power
(R∼ 2700).
4.1. Transmission line profile
One possible source of interference with the planetary He i ab-
sorption could be the Rossiter-McLaughlin effect (RME). How-
ever, this effect can be neglected as the stellar He i is very weak
in HD 209458. In fact, for similar targets where the stellar he-
lium is stronger, the RME is estimated to be smaller than 0.1%
(e.g., Nortmann et al. 2018; Salz et al. 2018). In addition, the
nearby Si i line at ∼10827 Å is almost nine times deeper than
the He i feature and leaves residuals in the average transmission
spectrum of 0.2 -- 0.4%. Thus, we estimated that the maximum
interference in the planetary He i absorption caused by the RME
should be smaller than 0.044% (i.e., 20 times smaller than the
measured transmission signal).
Around mid-transit, we noted a possible additional absorp-
tion feature about 10 -- 15 km s−1 blueward from the main absorp-
tion (Fig. 4). We measured in the HD 209458 spectra several ac-
tivity indicators related to Ca ii (infrared triplet), Na, and Hα us-
ing the program Starmod (Montes et al. 2000). There is no in-
dication that HD 209458 happened to be, at any moment during
the observations, in a higher state of activity than reported in
the literature (e.g., Czesla et al. 2017). Thus, a possible activity
impact on the transmission spectrum is unlikely to explain the
additional absorption.
Article number, page 5 of 7
Fig. 6. Light curve of the He i absorption centred on the observed core
of the line and using a width of ∆λ ∼0.9 Å (total ∆v = 25.3 km s−1).
Vertical dashed lines indicate the first, second, third, and fourth contact,
respectively. There is no evidence for any out of transit absorption.
Oklopci´c & Hirata (2018). For a given range of temperatures
and mass-loss rates ( M, where it refers to the total hydrogen
and helium mass loss), the radial density and velocity profiles
were computed by means of an isothermal Parker wind model
(Parker 1958). Afterwards, the continuity equations were solved
to derive the He 23S density profile. The He triplet absorption
was subsequently computed with the radiative transfer accord-
ing to the primary transit geometry (Ehrenreich et al. 2006).
The absorption coefficients and wavelengths for the three he-
lium metastable lines were taken from the NIST Atomic Spectra
Database3. Doppler line shapes were assumed at the temperature
of the helium model density. Additional broadening by the turbu-
lent velocity were not included as we found that most of the ab-
sorption comes from radii smaller than the Roche lobe (4.22 RP),
where turbulence is not expected to be important. A mean veloc-
ity of the gas along the line of sight (towards the observer) was
also included in order to account for a possible bulk motion of
the absorbing gas.
Figure 4 shows that the observed absorption can be well re-
produced for the helium triplet density obtained from that model
for a temperature of 6000 K and a mass-loss rate of 4.2×109 g s−1
(red curve). The mean molecular weight obtained from the
model for that fit is 0.76 amu. However, a degeneracy exists be-
tween the atmospheric temperature and M in the model, so that
M ranges between about 108 to 1011 g s−1 for temperatures of
4500 to 11500 K, respectively. With all due caution, these evap-
oration rates -- assuming a 90% H and 10% He atmosphere --
are similar to previous estimations of hydrogen escape based
on Lyα observations (e.g., Vidal-Madjar et al. 2003; Koskinen
et al. 2010; Bourrier & Lecavelier des Etangs 2013), or energy-
limited escape (Sanz-Forcada et al. 2011; Czesla et al. 2017),
which is reasonable as the atmosphere is assumed to be hydro-
gen dominated. In addition, the model shows that the weakest
component of the triplet is expected at around 1σ level, so its
non-detection is consistent with the given data quality. More de-
tails on the modelling and on the temperature and mass-loss rates
results will be given in a future paper.
3 https://www.nist.gov/pml/atomic-spectra-database
−0.04−0.020.000.020.04−1.0−0.50.00.51.0−0.04−0.020.000.020.04Orbital phase−1.0−0.50.00.51.0Average absorption level [%]−3−2−10123Time since mid−transit [h]A&A proofs: manuscript no. Alonso-Floriano_2019
We cannot exclude that this particular absorption compo-
nent is an artefact due to an unsatisfactory correction of an OH-
doublet of sky emission lines at ∼10830 Å (Fig. 2), but no such
effect is seen for the significantly brighter OH-emission line at
10832 Å. Alternatively, the presence of bad pixels in the area,
in particular on the blue side of the main He i component, could
also be responsible of the extra absorption. In addition to using
the standard bad pixel mask, we developed a method to identify
and correct bad pixels more carefully (see Sect. 3). It did not sig-
nificantly reduce the possible absorption feature on the blue side
of the main helium signal either.
If real, the signal can be fitted by an extra absorption compo-
nent in the average transmission spectrum from which a bulk ve-
locity shift of −13 km s−1 can be estimated (Fig. 4). This agrees
with the expected velocity of the escaping atmosphere at the
Roche lobe height (Salz et al. 2016). Although this suggests that
it originates at very large altitudes, which could be in the outer
layer of the thermosphere or even in the exosphere, a simple ra-
diative transfer model is still valid to estimate the velocity and
absorber amount of the feature and thus, tentatively fit the pro-
file of the absorption signal. However, this fit does not contain
information on the temporal variation of the feature that could
support its planetary origin. Further observations are needed to
confirm this additional absorption.
4.2. Stellar irradiation and He i signal
Nortmann et al. (2018) presented a relation between the strength
of the observed He i absorption and the stellar XUV (5 -- 504 Å)
irradiation, similar to that previously observed in stellar coronae
by Sanz-Forcada & Dupree (2008). The absorption line origi-
nates from neutral helium atoms in an excited metastable 23 S
state. The population of this level takes place after ionisation of
He i atoms by incoming irradiation from the host star, followed
by recombination in a cold environment. This radiation is gen-
erated in the corona and transition region of late type stars (late
F, G, K and M), and it is directly related to the level of activity,
which in turn depends mainly on stellar rotation. Thus close-in
gaseous planets around late type stars are prime targets to search
for the He triplet, considering also that active stars will likely
produce higher levels of XUV irradiation.
We place our new measurement of the He i absorption in the
transmission spectrum of HD 209458 b in context of this hy-
pothesis. In Fig. 7, we updated Fig. 4 presented in Nortmann
et al. (2018), which showed the empirical relation between
the stellar irradiation and the detectability of the He i signal.
The Y axis indicates the equivalent height of the helium sig-
nal (δRp) normalised by the atmospheric scale height (Heq). The
values were computed using the data provided by Nortmann
et al. (2018) and references therein, except for HD 209458 b
(Table 1, δRp/Heq = 46.9± 4.8), HAT-P-11 b (Allart et al. 2018;
Bakos et al. 2010; Deming et al. 2011, δRp/Heq = 103.4± 4.8),
and WASP-107 b (Allart et al. 2019; Anderson et al. 2017,
δRp/Heq = 87.7± 11.3). The X-ray and EUV (0.5 -- 50.4 nm) flux
used for the X axis are from Nortmann et al. (2018), except
for HD 209458 b. Its value (see Table 1) was calculated using
a modified version of the coronal model of Sanz-Forcada et
al. (2011) and applied as explained by Nortmann et al. (2018).
The coronal model was updated after an improved fit of the
summed XMM/EPIC spectrum (S/N = 3.2, log T (K) = 6.0 -- 6.3,
log EM (cm−3) = 49.52+0.22−0.48, see also Czesla et al. 2017) using
the same spectra as in Sanz-Forcada et al. (2011). The coronal
model was extended to cooler temperatures using UV line fluxes
Article number, page 6 of 7
Fig. 7. He i transmission signals currently detected (blue stars) and up-
per limits (black stars), as a function of the stellar irradiation below
504 Å at the planet distance. We show the equivalent height of the He i
atmosphere, δRp, normalised by the atmospheric scale height of the re-
spective planet's lower atmosphere, Heq.
from France et al. (2010). Because of the lack of X-ray infor-
mation for KELT-9 b, the corresponding value is indicated as a
lower limit. HD 209458 b has so far the weakest signal detected,
and it is also the planet that receives the least XUV flux from its
host star -- a relatively low-active G star. Therefore, this measure-
ment is in line with the suggested trend and relevant to anchor
the suspected activity relation at lower irradiation levels.
Although the distribution of all the signals in Fig. 7 is con-
sistent with a dependence on the XUV irradiation level, this may
not be the only factor. Oklopci´c (2019) modelled the strength of
He i absorption in irradiated planetary atmospheres depending
on the spectral type of the hosting star. They suggested that the
ratio between extreme- and mid-UV irradiation fluxes determine
the amplitude of this absorption, which particularly favours K
stars. It remains to be investigated to what extent the mid-UV
flux of HD 209458 is responsible for the weaker helium absorp-
tion compared to the other four detections in planets orbiting K
stars.
5. Conclusions
We present a solid detection of He i λ 10830 Å in the trans-
mission spectrum of the hot Jupiter HD 209458 b at a level of
0.91 ± 0.10 %. It concludes a search for He i in this planet atmo-
sphere for over a decade (Seager & Sasselov 2000; Moutou et
al. 2003; Nortmann et al. 2018). The strength of the detection is
consistent with the empirical relationship proposed between the
helium signal and the host star activity (Nortmann et al. 2018).
We tentatively detect additional absorption on the blue side
of the main helium signal at about −13 km s−1. Although this
could be due to cometary-tail like escape, we are not confident
yet about the reliability of the feature. In addition, there is no
evidence for pre- or post-transit absorption in the He i triplet.
Our spectral detection is consistent with models of atmospheric
escape with total hydrogen and helium mass-loss rates of 108 --
1011 g s−1 depending on the assumed temperature of the upper
atmosphere.
Acknowledgements. We thank P. Molliere and A. Wyttenbach for the nice sci-
entific discussions during the preparation of this publication. F.J.A.-F. and I.S.
acknowledge funding from the European Research Council (ERC) under the Eu-
0.11.010.0100.00204060801001200.11.010.0100.0FluxXEUV [Wm−2]020406080100120δRp/HeqWASP−69 bHD 189733 bHD 209458 bKELT−9 bHAT−P−11 bWASP−107 bGJ 436 bF. J. Alonso-Floriano et al.: He i in the transmission spectrum of HD 209458 b
ropean Union Horizon 2020 research and innovation programme under grant
agreement No 694513. CARMENES is funded by the German Max-Planck-
Gesellschaft (MPG), the Spanish Consejo Superior de Investigaciones Cientí-
ficas (CSIC), the European Union through FEDER/ERF FICTS-2011-02 funds,
and the members of the CARMENES Consortium (MaxPlanck-Institut für As-
tronomie, Instituto de Astrofísica de Andalucía, Landessternwarte Königstuhl,
Institut de Ciències de l'Espai, Insitut für Astrophysik Göttingen, Universidad
Complutense de Madrid, Thüringer Landessternwarte Tautenburg, Instituto de
Astrofísica de Canarias, Hamburger Sternwarte, Centro de Astrobiología and
Centro Astronómico Hispano-Alemán), with additional contributions by the
Spanish Ministry of Economy, the German Science Foundation through the Ma-
jor Research Instrumentation Programme and DFG Research Unit FOR2544
"Blue Planets around Red Stars", the Klaus Tschira Stiftung, the states of Baden-
Württemberg and Niedersachsen, and by the Junta de Andalucía. Financial sup-
port was also provided by the Universidad Complutense de Madrid, the Co-
munidad Autónoma de Madrid, the Spanish Ministerios de Ciencia e Inno-
vación and of Economía y Competitividad, the State Agency for Research of the
Spanish MCIU through the "Center of Excellence Severo Ochoa" and Science
& Technology Facility Council Consolidated, and the Fondo Social Europeo.
The corresponding funding grants are: ESP2014 -- 54362 -- P, ESP2014 -- 54062 --
R, AYA2015-69350 -- C3 -- 2 -- P, BES -- 2015 -- 074542, AYA2016-79425 -- C3 -- 1/2/3 --
P, ESP2016 -- 76076 -- R, ESP2017 -- 87143 -- R, SEV -- 2017 -- 0709, ST/P000592/1.
Based on observations collected at the Centro Astronómico Hispano Alemán
(CAHA) at Calar Alto, operated jointly by the Max -- Planck Institut für As-
tronomie and the Instituto de Astrofísica de Andalucía. We thank the anonymous
referee for their insightful comments, which contributed to improve the quality
of the manuscript.
A147
L45
568, 377
References
Allart, R., Bourrier, V., Lovis, C., et al. 2018, Science, 362, 1384
Allart, R., Bourrier, V., Lovis, C., et al. 2019, A&A, 623, A58
Anderson, D. R., Collier Cameron, A., Delrez, L., et al. 2017, A&A, 604, A110
Bakos, G. Á., Torres, G., Pál, A., et al. 2010, ApJ, 710, 1724
Ben-Jaffel, L. 2007, ApJ, 671, L61
Bourrier, V., & Lecavelier des Etangs, A. 2013, A&A, 557, A124
Bourrier, V., Ehrenreich, D., Wheatley, P. J., et al. 2017, A&A, 599, L3
Bourrier, V., Lecavelier des Etangs, A., Ehrenreich, D., et al. 2018, A&A, 620,
Brogi, M., Line, M., Bean, J., Désert, J.-M., & Schwarz, H. 2017, ApJ, 839, L2
Caballero, J. A., Guàrdia, J., López del Fresno, M., et al. 2016, SPIE, 9910, 0E
Charbonneau, D., Brown, T. M., Latham, D. W., & Mayor, M. 2000, ApJ, 529,
Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L. 2002, ApJ,
Crossfield, I. J. M., Knutson, H., Fortney, J., et al. 2012, ApJ, 752, 81
Czesla, S., Salz, M., Schneider, P. C., Mittag, M., & Schmitt, J. H. M. M. 2017,
A&A, 607, A101
Deming, D., Seager, S., Richardson, L. J., & Harrington, J. 2005, Nature, 434,
Deming, D., Brown, T. M., Charbonneau, D., Harrington, J., & Richardson, L. J.
2005, ApJ, 622, 1149
Deming, D., Sada, P. V., Jackson, B., et al. 2011, ApJ, 740, 33
Dorner, B., Giardino, G., Ferruit, P., et al. 2016, A&A, 592, A113
Ehrenreich, D., Tinetti, G., Lecavelier des Etangs, A., Vidal-Madjar, A., & Selsis,
F. 2006, A&A, 448, 379
Ehrenreich, D., Lecavelier des Etangs, A., Hébrard, G., et al. 2008, A&A, 483,
Ehrenreich, D., Bourrier, V., Wheatley, P. J., et al. 2015, Nature, 522, 459
France, K., Stocke, J. T., Yang, H., et al. 2010, ApJ, 712, 1277
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018, A&A, 616, A1
Hawker, G. A., Madhusudhan, N., Cabot, S. H. C., & Gandhi, S. 2018, ApJ, 863,
Henry, G. W., Marcy, G. W., Butler, R. P., & Vogt, S. S. 2000, ApJ, 529, L41
Høg, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A, 355, L27.
Hoeijmakers, H. J., de Kok, R. J., Snellen, I. A. G., et al. 2015, A&A, 575, A20
Kausch, W., Noll, S., Smette, A., et al. 2015, A&A, 576, A78
Knutson, H. A., Charbonneau, D., Noyes, R. W., Brown, T. M., & Gilliland, R. L.
Knutson, H. A., Charbonneau, D., Allen, L. E., Burrows, A., & Megeath, S. T.
2007, ApJ, 655, 564
2008, ApJ, 673, 526
Koskinen, T. T., Yelle, R. V., Lavvas, P., & Lewis, N. K. 2010, ApJ, 723, 116
Kulow, J. R., France, K., Linsky, J., & Loyd, R. O. P. 2014, ApJ, 786, 132
Lammer, H., Kasting, J. F., Chassefière, E., et al. 2008, Space Sci. Rev., 139, 399
Lecavelier des Etangs, A., Ehrenreich, D., Vidal-Madjar, A., et al. 2010, A&A,
514, A72
Lavie, B., Ehrenreich, D., Bourrier, V., et al. 2017, A&A, 605, L7
740
933
L11
Mazeh, T., Naef, D., Torres, G., et al. 2000, ApJ, 532, L55
Mansfield, M., Bean, J. L., Oklopci´c, A., et al. 2018, ApJ, 868, L34
Montes, D., Fernández-Figueroa, M. J., De Castro, E., et al. 2000, A&AS, 146,
103
405, 341
Moutou, C., Coustenis, A., Schneider, J., Queloz, D., & Mayor, M. 2003, A&A,
Nortmann, L., Pallé, E., Salz, M., et al. 2018, Science, 362, 1388
Oklopci´c, A., & Hirata, C. M. 2018, ApJ, 855, L11
Oklopci´c, A. 2019, ApJ, submitted, arXiv:1903.02576
Parker, E. N. 1958, ApJ, 128, 664
Quirrenbach, A., Amado, P. J., Caballero, J. A., et al. 2016, SPIE, 9908, 12
Quirrenbach, A., Amado, P. J., Ribas, I., et al. 2018, SPIE, 10702, 0W
Salz, M., Czesla, S., Schneider, P. C., & Schmitt, J. H. M. M. 2016, A&A, 586,
A75
A111
Salz, M., Czesla, S., Schneider, P. C., et al. 2018, A&A, 620, A97
Sanz-Forcada, J., & Dupree, A. K. 2008, A&A, 488, 715
Sanz-Forcada, J., Micela, G., Ribas, I., et al. 2011, A&A, 532, A6
Schwarz, H., Brogi, M., de Kok, R., Birkby, J., & Snellen, I. 2015, A&A, 576,
Seager, S., & Sasselov, D. D. 2000, ApJ, 537, 916
Seifert, W., Sánchez Carrasco, M. A., Xu, W., et al. 2012, SPIE, 8446, 33
Shulyak, D., Reiners, A., Nagel, E., et al. 2019, A&A, 626, A86
Smette, A., Sana, H., Noll, S., et al. 2015, A&A, 576, A77
Snellen, I. A. G., Albrecht, S., de Mooij, E. J. W., & Le Poole, R. S. 2008, A&A,
Snellen, I. A. G., de Kok, R. J., de Mooij, E. J. W., & Albrecht, S. 2010, Nature,
Spake, J. J., Sing, D. K., Evans, T. M., et al. 2018, Nature, 557, 68
Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163
Swain, M. R., Tinetti, G., Vasisht, G., et al. 2009, ApJ, 704, 1616
Torres, G., Winn, J. N., & Holman, M. J. 2008, ApJ, 677, 1324
Vidal-Madjar, A., Lecavelier des Etangs, A., Désert, J.-M., et al. 2003, Nature,
Vidal-Madjar, A., Désert, J.-M., Lecavelier des Etangs, A., et al. 2004, ApJ, 604,
487, 357
465, 1049
422, 143
L69
Vidal-Madjar, A., Huitson, C. M., Bourrier, V., et al. 2013, A&A, 560, A54
Wright, J. T., Fakhouri, O., Marcy, G. W., et al. 2011, PASP, 123, 412
Zechmeister, M., Anglada-Escudé, G., & Reiners, A. 2014, A&A, 561, A59
Article number, page 7 of 7
|
1204.2014 | 1 | 1204 | 2012-04-10T00:17:25 | Star Hoppers: Planet Instability and Capture in Evolving Binary Systems | [
"astro-ph.EP",
"astro-ph.SR"
] | Many planets are observed in stellar binary systems, and their frequency may be comparable to that of planetary systems around single stars. Binary stellar evolution in such systems influences the dynamical evolution of the resident planets. Here we study the evolution of a single planet orbiting one star in an evolving binary system. We find that stellar evolution can trigger dynamical instabilities that drive planets into chaotic orbits. This instability leads to planet-star collisions, exchange of the planet between the binary stars ("star-hoppers"), and ejection of the planet from the system. The means by which planets can be recaptured is similar to the pull-down capture mechanism for irregular solar system satellites. Because planets often suffer close encounters with the primary on the asymptotic giant branch, captures during a collision with the stellar envelope are also possible. Such capture could populate the habitable zone around white dwarfs. | astro-ph.EP | astro-ph |
Draft version November 15, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
STAR HOPPERS: PLANET INSTABILITY AND CAPTURE IN EVOLVING BINARY SYSTEMS
Kaitlin M. Kratter & Hagai B. Perets
Institute for Theory and Computation, Harvard-Smithsonian Center for Astrophysics, 60 Garden St.; Cambridge, MA, USA 02138
Draft version November 15, 2018
ABSTRACT
Many planets are observed in stellar binary systems, and their frequency may be comparable to
that of planetary systems around single stars. Binary stellar evolution in such systems influences the
dynamical evolution of the resident planets. Here we study the evolution of a single planet orbiting one
star in an evolving binary system. We find that stellar evolution can trigger dynamical instabilities
that drive planets into chaotic orbits. This instability leads to planet-star collisions, exchange of
the planet between the binary stars ("star-hoppers"), and ejection of the planet from the system.
The means by which planets can be recaptured is similar to the pull-down capture mechanism for
irregular solar system satellites. Because planets often suffer close encounters with the primary on the
asymptotic giant branch, captures during a collision with the stellar envelope are also possible. Such
capture could populate the habitable zone around white dwarfs.
1.
INTRODUCTION
Most of the (> 2000) planets and planet candidates
discovered to date are hosted by single stars (Howard
et al. 2011). This bias towards single stars may be pri-
marily due to the avoidance of multiple systems in target
selection: the limited number of planet discoveries in bi-
nary systems suggests that the planet frequency in mul-
tiples may be comparable to that of single stars (Eggen-
berger et al. 2004; Bonavita & Desidera 2007; Mugrauer
& Neuhauser 2009). Given the frequency of binary sys-
tems (Raghavan et al. 2010), a large fraction of all exo-
planetary systems may reside in binaries. Here we focus
on the dynamics of planets in evolving binary systems
near the peak of the binary separation distribution: from
∼ 50 − 100 AU.
Though various studies have explored the possible for-
mation and dynamical evolution of planets in binary
systems (Holman & Wiegert 1999; David et al. 2003;
Mudryk & Wu 2006; Perets 2010; Moeckel & Veras 2012),
the effects of stellar evolution on such systems have not
been thoroughly explored. The influence of stellar evo-
lution on planets in singleton systems has been studied
by Debes & Sigurdsson (2002); Villaver & Livio (2009);
Veras et al. (2011). Perets (2010) described qualitatively
the evolution of planets in evolved binary systems, in-
cluding both circumstellar and circumbinary planets. Ve-
ras & Tout (2012) have recently conducted a quantitative
study of circumbinary planetary dynamics in evolving
systems.
In this paper we focus on the effects of binary stel-
lar evolution on planetary systems in "satellite" orbits,
those around one star in the system. We consider the
influence of slow mass loss via stellar winds, and demon-
strate that these systems have rich dynamical behavior,
including planetary ejections, "star hopping" (exchanges
of planets between the binary stars), capture of planets
into stable orbits, collisions with either of the stars, and
interactions potentially leading to capture into close or-
bits. A similar mechanism that operates in triple stellar
systems, dubbed the Triple Evolution Dynamical Insta-
bility (TEDI) has been recently explored by Perets &
Kratter (2012).
In the following we explore the outcomes of this insta-
bility using few-body simulations, and provide a quali-
tative and semi-quantitative understanding of the prop-
erties of such systems. We also provide an estimate for
the fraction of systems that are subject to the planetary
version of the TEDI, given various assumptions for the
population of planetary systems in binaries.
We begin in Section 2 by describing the orbital evolu-
tion and stability criterion for planets in evolving bina-
ries. We next explain the capture mechanism for planets
around the secondary star in the context of the circular
restricted three body problem (Section 3). In Section 4
we describe a series of n-body experiments that we use
to explore the outcomes of the chaotic planetary orbits
induced by stellar mass loss.
In Section 5 we present
probabilities for collision and capture for planets in our
numerical experiments, and provide a summary of ex-
pected outcomes. In Section 6 we explore the frequency
of unstable planetary binary systems. Finally, in Section
7 we conclude with several observational predictions for
next generation surveys.
2. DYNAMICAL EVOLUTION OF CIRCUMSTELLAR
PLANETS IN EVOLVING BINARY SYSTEMS
In order to remain dynamically stable, a triple system
of gravitating objects requires a hierarchical configura-
tion with orbits that are well separated. For satellite type
planetary-binary systems, there is a critical ratio between
the semi-major axes of the binary and the planet beyond
which the system is unstable. The stability boundary
is quite complex in detail, set by the overlap of mean
motion resonances (Wisdom 1980; Mudryk & Wu 2006),
however it can be well fit as a function of binary mass
ratio and eccentricity (Holman & Wiegert 1999). For cir-
cular binary orbits, which we consider here, the radius of
the stable region around the primary is:
(cid:18)
(cid:19)
Rp =
0.464 − 0.38
m2
m1 + m2
a∗
(1)
where a∗ is the stellar semi-major axis, and m1 and m2
are the masses of the primary and secondary, respec-
tively. Note that this is identical to Equation (1) of Hol-
2
man & Wiegert (1999), but without the quadratic de-
pendence on eccentricity. These authors find that plan-
ets with semi-major axis ap > Rp are destabilized within
104 binary orbital periods, (see also the results of David
et al. 2003 and Fatuzzo et al. 2006, who calculate typical
instability timescales). The size of the stability region
around the secondary, Rs, is obtained by the substitu-
tion m1 ↔ m2. Note that the initial eccentricity of the
planet, which we set to zero here, is unimportant for sta-
bility considerations (Mudryk & Wu 2006).
We now consider the impact of slow (adiabatic) and
isotropic mass loss from the primary star in the binary,
where the mass loss timescale is long compared to the
planetary and stellar orbital periods. Mass loss is the
byproduct of stellar evolution and occurs through stellar
winds. Adiabatic mass loss drives the orbits to larger
semi-major axes:
af =
Mi
Mf
ai,
(2)
(Hadjidemetriou 1963), where Mf is the final mass of
the system after its evolution, and Mi and ai are the
initial mass and initial semi-major axis of the system,
respectively. This relation derives from a perturbation
analysis that shows that the the specific angular mo-
mentum GM a(1 − e2) (per reduced unit mass) is con-
served, and the assumption that mass loss is slow and
isotropic, which guarantees that eccentricity is also con-
served. These assumptions are only accurate for planets
far from their host star (i.e. beyond the tidal effects of
interaction with the expanding envelope of the evolving
primary), and where mass transfer is negligible. The
outcomes of the latter possibility have been discussed
in the context of planets orbiting a single evolving star
(e.g. Livio & Soker 1984; Villaver & Livio 2007; Nord-
haus et al. 2010 and references therein).
Although the orbits of both the stellar companion and
the planet expand due to mass loss, the change to the
planet's orbit is much larger than the change in the bi-
nary's orbit because the mass loss within the planet-star
system comprises a greater fraction of the enclosed mass.
Allowing the planet mass to go to zero, the ratio between
the semi-major axis of the planet to stellar companion
therefore increases by a factor of:
(cid:18) m1,i
(cid:19)(cid:18) m1,f + m2
(cid:19)
(cid:18) ap,f
(cid:19)
a∗,f
(cid:18) ap,i
(cid:19)
a∗,i
/
=
m1,f
m1,i + m2
.
(3)
where the subscripts correspond to the system compo-
nents (1, 2, and p for the evolving star, its stellar com-
panion and its planetary companion, respectively), and i
and f corresponding to the initial and the final system,
respectively. Relative orbital expansion due to mass loss
can shift planets across the stability threshold.
Combining Equations (1) and (3), we can define the
approximate stability boundary as a function of the evo-
lutionary state (i.e. current primary mass) of a given
binary system. In Figure 1, we show the theoretical evo-
lution of several planetary orbits as a function of their
separation ratio for the evolution of a 2M(cid:12) star in a bi-
nary orbit with a 1M(cid:12) companion. In this system, the
ratio of the planetary to stellar orbit, ap/a∗, expands by
nearly a factor of two as the primary evolves to become a
white dwarf (hereafter WD). At the same time, the criti-
Fig. 1. -- Top panel: Evolution of the ratio of planet to stellar
semi-major axis (thin colored lines) and critical stability bound-
ary (thick dashed line) for a mass losing 2M(cid:12) with solar mass
companion. Initial semi-major axis ratios which cross the stability
threshold will enter a chaotic phase of evolution. We show only
the ∼ 2Myr phase where mass loss is most rapid. The stability
threshold declines more sharply than the mass because the stability
boundary also shrinks due to the change in µ = m2/(m2 + m1).
Bottom panel: the corresponding stellar mass and stellar radius
evolution for the evolving primary.
cal boundary for stability around the primary shrinks as
the primary becomes less massive than its companion.
We show the evolving mass and radius of the primary as
well: the stellar model is calculated from the Single Stel-
lar Evolution (SSE) code of Hurley et al. (2000), which
evolves the primary from m1 = 2.0 → 0.55M(cid:12).
Planets with semi-major axis ratios that exceed the
critical point in Equation (1) enter into a phase of chaotic
evolution. Note that we use chaotic in the technical
sense, wherein planets with nearly identical initial con-
ditions undergo dramatically different orbital evolution
See Mudryk & Wu (2006) for a discussion of measured
Lyapunov exponents for the problem without mass loss.
After becoming unstable, planets can suffer close en-
counters with either component of the binary. Intrigu-
ingly, the instability can also lead to the long-term cap-
ture of planets around the secondary star. We now dis-
cuss the capture mechanism in more detail.
3. CAPTURE AND THE JACOBI CONSTANT
Before describing our numerical results, we begin with
a brief description of the capture process, in which a
planet originally orbiting the primary star is destabilized
due to stellar mass loss, and then recaptured into a stable
orbit around the companion. We do so by considering
the circular restricted three-body problem: coplanar test
particles moving about two massive bodies on a circular
orbit.
Capture occurs through a process reminiscent of the
suggested pull down mechanism for the capture of
the Moon (Ruskol 1961) and irregular satellite capture
around the gas giants in the Solar System (Heppenheimer
& Porco 1977; Astakhov et al. 2003). In these cases, the
body that captures a satellite does so via gaining mass.
We show here that mass loss by the primary can also
lead to capture.
Capture is controlled by the evolution of a particle's
Jacobi constant, CJ . Recall that the Jacobi constant
is the only conserved integral of motion in the circular
restricted three body problem (hereafter CR3BP):
CJ = 2
+ n2(x2 + y2) − (v2
x + v2
y + v2
z ) (4)
(cid:18) µ1
r1
(cid:19)
+
µ2
r2
where µ1 and µ2 are the mass ratios of the stars, r1 and
r2 are the separations of the particle from each star, n is
the mean motion, and the cartesian components are the
positions and velocities of the particle in the non-rotating
reference frame (Murray & Dermott 2000). A particle's
CJ defines the parameter space in which it may orbit the
binary. The boundaries of the permitted region can be
visualized through the use of zero-velocity curves. These
contours are obtained by setting the planet velocities to
zero and solving Equation (4) for the positions in the
non-rotating reference frame.
By adding mass loss into the CR3BP, CJ is no
longer conserved, because µ evolves in time. However,
Luk'yanov (2009) has shown that this so-called quasi-
Jacobi integral is mathematically well behaved in the
case of mass transfer in a binary system, suggesting that
it still specifies the bounds of the planet's orbit at any
instant.
The evolution of CJ due to mass loss dictates whether a
given particle can pass between the stars, or must remain
in orbit about one or the other. Figure 2 shows the zero
velocity curves for a planet at four epochs in the evolution
of a system with m1 = 2.0 → 0.55M(cid:12), m2 = 1.0M(cid:12),
and initial semi-major axis, a∗,i = 90 AU. Figure 3 is
complementary, and shows the evolution of the planetary
and stellar separations, as well as the particle's CJ as a
function of time.
The fall and rise of CJ shown in both Figures 2 and 3 is
characteristic of host switching. As shown in Figure 3, we
can estimate the evolution of CJ analytically for a par-
ticle's orbit that expands with mass loss as dictated by
Equation (2). At early times, we keep the planet fixed at
its Keplerian rotation velocity. Once the planet nears the
instability threshold, the orbit is highly non-Keplerian.
In this regime we interpolate between CJ calculated for
two assumptions: a) that the planet remains on a Ke-
plerian orbit about the primary, and b) that the planet
switches to a Keplerian orbit at the same separation,
but about the secondary. The disagreement between the
analytic estimate and the actual particle value is worst
where orbits depart most from Keplerian (i.e.
in the
planet hopping region). The relatively better agreement
3
at late times suggests that the particle is well described
by a Keplerian orbit, and that the orbital energy is con-
sistent with that of the initial orbit.
In the case shown here (the simulations are described
in the following section), the planet's CJ at first pro-
hibits transfer between the two stars, then permits it,
and then once again forbids it. If a planet is in the sta-
ble region around the secondary when the curves close
off, the planet becomes trapped. One can distinguish
between open and closed zero-velocity curves by com-
paring the particles' Jacobi constants to the value at the
first Lagrange point ( L1), or:
CJ (L1) ≈ 3 + 3(4/3)µ(2/3)
2
− 10µ2/3
(5)
(Murray & Dermott 2000).
Not all systems undergo this complete cycle. For some,
when mass loss stops, the zero velocity curves remain
open, and capture is less likely. Note that the stability
threshold defined in Equation (1) is crossed only after the
curves have opened: planets can remain on stable orbits
in the face of open zero-velocity curves.
We show in Section 5 that the relative value of the the
average Jacobi constant of the surviving particles and at
the L1 point, ¯CJ−CJ (L1), correlates with the probability
of planets being captured.
4. FEW-BODY SIMULATIONS
We use few-body simulations to explore the evolution
of planets in evolving binary systems. We focus on a lim-
ited phase space, which allows us to explore the complex
dynamical behavior as a function of system parameters
and mass loss rates. We consider only point masses, and
limit ourselves to circular binaries hosting initially circu-
lar, coplanar planets, and a single primary mass and evo-
lutionary pathway. Initial planet eccentricity is unlikely
to be important, as shown by Mudryk & Wu (2006).
We use a fourth order Hermite integrator based on the
algorithm of Hut et al. (1995), with an additional mass
loss term. Mass loss rates and evolving stellar radii are
calculated from the SSE models of Hurley et al. (2000).
We use the default values for mass loss rates (assum-
ing solar metallicity) provided in the publicly available
version of the code. Each run consists of two massive
bodies, and 100 test particles representing planets. The
time step for all particles is set to 10−2 − 10−3 of the en-
counter timescale for the closest pair of particles in the
simulation, where the smaller step is required for good
energy conservation for the fastest mass loss rates1. The
stellar mass is updated on the same time step as the
test particles, and all particles share the same time step.
When mass loss is included, energy is not conserved by
default, however in test runs with no mass loss, the code
conserves energy to 1 part in 1012 over a 5 Myr run. We
also confirm that particles on both circular and chaotic
orbits (e.g. bouncing between the stars) conserve their
Jacobi constants over a 5 Myr run (see Moeckel & Veras
2012). We check that the evolution of the stars and test
particles follows the expected orbital expansion ( Equa-
tion 2) up until the point of instability. Note that for the
realistic (fast) mass loss case, a small eccentricity is intro-
1 The outcomes remain the same under an order of magnitude
reduction in the time step.
4
Fig. 2. -- Evolution of the the zero-velocity curves for a planet as the primary (cyan) loses mass. The white regions indicate where the
planet may orbit. The bottleneck surrounding the first Lagrange point opens midway through the mass loss phase, allowing the planet to
orbit both components of the binary. The dashed-tail shows an episode of orbital bouncing, which continues while the curves remain open.
By the end of the mass loss phase, the zero-velocity surface has been pinched off, and the planet is trapped around the secondary. A movie
of this evolution is available at http://www.cfa.harvard.edu/kkratter/BinaryPlanets/zvelmovie.mpg.
duced into the stellar binary -- this results in oscillations
in the Jacobi constant for some surviving test particles.
4.1. Simulation Details
We simulate evolving binaries containing primaries
which evolve from M = 2 → 0.55M(cid:12). Such A stars host
planets (Johnson et al. 2007), undergo extensive mass
loss in a Hubble time and are frequently found in multiple
systems (Raghavan et al. 2010). We consider secondary
masses ranging from m2 = 0.5−1.7M(cid:12), and initial binary
separations from a∗,i = 75 − 105 AU. We fix the planet
separation at 15 AU: far enough to avoid any interaction
with the primary's envelope, and close enough to be well
inside the (initial) stability region around the host star.
Test particles are initialized with zero eccentricity, ran-
dom phase, on coplanar orbits. In order to expedite the
calculations, we evolve the planetary and stellar orbits
with the analytic formulae for the first ∼ 1490 Myr, un-
til the star reaches the beginning of the asymptotic giant
branch (AGB), at which point the mass loss rate rapidly
increases (shown in Figure 1).
We consider only the evolution of the more massive
star, which leaves the main sequence first (the planet
host). In Section 5 we discuss possible outcomes when
the second star evolves off of the main sequence.
In total we explore 70 different binary configurations.
For each binary configuration, we also explore the influ-
ence of mass loss rate and collisional cross section. Even
varying only two parameters produces a complex set of
outcomes because the probabilistic fate of a planet de-
pends not only on the initial and final configurations, but
also on the rate at which they pass through each state,
which is determined by the mass loss rate and expansion
of the primary's atmosphere. We run our fiducial cal-
culations for 30 Myr following the end of the mass loss
phase, or 104 orbits, as most planets that survive to 10
Myr remain out to 100 Myr in several test cases. As seen
by Rabl & Dvorak (1988) and Holman & Wiegert (1999),
in the case with no mass loss, most of the orbital evolu-
tion occurs on timescales of less than 300 binary orbits.
We discuss these timescales further in Section 5.3.
Even though all the particles begin with nearly identi-
cal CJ (small variations are introduced due to variations
in the planets' initial phases), mass loss induces a spread
in CJ because they undergo different chaotic evolution-
ary paths while mass loss is ongoing. This spread pro-
duces a range of outcomes for planets in a given system.
During the course of the simulation we keep track of
each particles closest approach to either star. Particles
which penetrate the specified stellar radii are removed
from the calculation and identified as collisions. For the
primary, the collisional radius is set to be the maximum
of either the stellar radius calculated from Hurley et al.
(2000) or the star's approximate tidal radius as a WD:
3.8×10−3 AU. The secondary's radius is fixed at 1R(cid:12) for
all masses to simplify the interpretation of probabilities.
5
1 in the Appendix. In the figure, each panel shows the
outcome for a fixed initial binary separation for a range
of masses. For all systems that become unstable, the
most likely outcome for a planet is collision with one of
the stars.
Several trends are readily observed. Collisions typi-
cally increase towards higher masses at fixed separation,
while tidal encounters decrease. Both trends are the re-
sult of the instability threshold being reached at an ear-
lier evolutionary state, when the primary has an inflated
radius; wider encounters become collisions, and encoun-
ters within a few stellar radii are prohibited. Note that in
order to collide with either star, planets must be set onto
low angular momentum orbits. As a result, collisions in-
volve at least one passage back and forth between the
stars.
At larger separations and stellar masses, planets are
frequently (10%) captured by the secondary star. Cap-
tured particles come in two varieties: those whose en-
ergies as measured via CJ dictate that they remain in
orbit only about the secondary, and those who in prin-
cipal may pass back into orbit around the primary, but
show no evidence of such movement over the course of
the simulation. These two are distinguished in Figure 4
by the sign of ¯CJ − CJ (L1), where only those with posi-
tive values are truly trapped (closed zero velocity curves).
Clearly, some with open curves (negative values) do sur-
vive for the duration of the simulation. This result is in
line with the findings of Heppenheimer & Porco (1977).
They attribute this behavior to planets entering into spe-
cial orbits that avoid the conjunction of apocenter with
the L1 point. In general, more planets are captured in
systems whose zero-velocity curves close at the end of
mass loss.
Both classes of captured planets typically have pro-
grade, eccentric orbits (e > 0.4), and rapid, retrograde
precession in the frame of the binary. Using the method
of the surface of section, we identify many of these plan-
ets as having quasi-periodic, though formally chaotic or-
bital parameters (Murray & Dermott 2000). Their long
term (> 100Myr) survival remains uncertain. Neverthe-
less, weaker tidal effects may also dissipate energy, caus-
ing such orbits to shrink and circularize (Moeckel & Veras
2012).
The captured planets typically have orbits which graze
the stability boundary, as shown in Figure 3. Such plan-
ets are extremely likely to go unstable when the second
star in the system evolves to become a WD. A secondary
instability cycle might re-populate the original host with
planets on eccentric orbits.
Close approaches, q, within 1R∗ < q < 3R∗, are also
common ( 10%), we discuss our treatment of these plan-
ets, and the plausible results of close encounters and
physical collisions below.
5.2. Tidal encounters
By excluding planets with previous close approaches
(q <∼ 3R(cid:12)) from our calculated probabilities for both cap-
ture and collision, we expect that our neglect of tidal ef-
fects is unlikely to bias our results. When particle orbits
are chaotic, they do not evolve smoothly to ever closer
values of q, but rather make order of magnitude jumps
in separation from one close approach to the next. To
Fig. 3. -- Time evolution of the distance and Jacobi constant for
the same system shown in Figure 2, where one planet has become
trapped in orbit about the secondary. The vertical line shows when
the particle crosses the stability threshold defined in Equation (1).
The thick black line indicates the stellar separation, the purple
line indicates the planet separation from the primary, and the pink
line the distance from the secondary. The thick dashed red and
grey lines indicate the stability boundaries around the primary
and secondary, Rp and Rs respectively. The thin black line shows
the initial analytic evolution predicted for the planetary orbit. The
top panel shows the particle's CJ (green), the value of CJ at the
L1 point (black), and the analytic value for CJ of the particle
on a Keplerian orbit (blue dashed), with separation specified by
Equation (2) as discussed in Section 3.
Because we neglect tidal effects, we also identify any
particle that has suffered a close approach within 3 tidal
radii of either star (but not a direct collision) as having
had a tidal interaction. We define the tidal radius to be:
(cid:18) M(cid:63)
(cid:19)1/3
rt ≈
Mp
rp.
(6)
Particles that do not suffer tidal interactions or colli-
sions are identified as bound to either the primary or
secondary, although typically no planets remain around
the primary2.
5. RESULTS
We begin by showing the probabilities of capture, colli-
sion, and tidal encounters for the fiducial runs with real-
istic mass loss rates and stellar radii in Figure 4. We then
describe in turn the importance of the mass loss rate, and
physical radius in controlling captures and collisions.
5.1. Outcomes from fiducial few-body simulations
Figure 4 shows the probabilities for captures, collisions,
and tidal interactions for 60 of our 70 binary systems.
The tabulated results for all systems are shown in Table
2 As shown in Table 1, there are five systems which retain planets
in one case (m2 = 0.8M(cid:12), a∗,i = 105AU),
around the primary:
the system evolves up to the edge of the stability threshold, and
because of phase variations only some planets become unstable. We
repeated this run several times and found similar results. For the
remaining systems, those classified as bound to the primary are on
bouncing orbits that we expect will ultimately result in collisions:
they happen to be in orbit about the primary at the conclusion of
the simulation.
1.21.41.61.82.02.22.4Time (Myrs)50100150200AUa∗RpRs~rm1−~rp~rm2−~rp3.84.24.65.05.4CJm2=1.0, a∗,i=90CJCJ(L1)CJ analytic6
Fig. 4. -- Probabilities for collision, capture, and tidal encounters in our fiducial models. Masses are in units of M(cid:12) and separations in
AU. Each panel represents a different semi-major axis for the binary, while different secondary masses are shown along the x-axis of each
panel. All collisions (black), and collisions with the primary (red dashed) typically increase with increasing mass of the secondary, and
decrease with increasing binary separation. The decrease in tidal encounters (cyan plusses) with increasing secondary mass is due to more
collisions with the AGB atmosphere due to the earlier onset of chaotic orbits. Whether or not the zero-velocity curves have pinched off,
preventing continued planet bouncing, is indicated by the sign of ¯CJ − CJ (L1) (blue diamonds). Positive values show that the curves are
closed.
demonstrate this behavior, we show in Figure 5 the evolu-
tion of the closest approach distance for a small sample
of particles from one run. We see that a particle that
reaches a separation of a few stellar radii has likely not
had previous encounters at comparable distances, and
thus not experienced significant tidal damping. For plan-
ets that suffer encounters within a few stellar radii, tidal
disruption is the most likely outcome.
5.2.1. Tidal disruption versus capture
Planet encounters at many stellar radii lose almost no
energy because the energy dissipated scales as a−6
p , where
ap is the periastron. When tidal effects become impor-
tant at a few tidal radii, close approaches sap energy
from the planet's orbit.
In principal these encounters
can decrease the planet's velocity (and thus increase CJ ),
enough to bring it from an unstable to stable orbit in a
binary (or from unbound to bound in a single star sys-
tem). Tidal capture has been suggested as a formation
mechanism for X-ray binaries through close encounters
between stars and compact objects in globular clusters
(Fabian et al. 1975), and for close-in planets (Faber et al.
2005).
Recent hydrodynamical simulations suggest that close
approaches lead to the disruption or ejection of a Jupiter-
mass planet, rather than long-term survival (Guillochon
Fig. 5. -- Closest approach q in units of solar radii for a sample of
test particles in one binary system. Each color represents a differ-
ent particle trajectory. Note that particles typically undergo large
jumps in close approach, suggesting that their orbits are chaoti-
cally sampling the orbital parameter space, not decaying to ever
smaller separations.
et al. 2011). At relatively larger distances, energy is dis-
sipated through tidal heating of the planet, causing its
orbit to shrink. As the orbit decays, the planet begins
losing mass through tidal stripping; this stripping and
mass transfer typically lead to the ejection of the planet
to larger separations at high velocity, or to total plane-
tary disruption.
The binary scenario is somewhat different than the
case of a single-star planet interaction. Tidally stripped
planets that would have been ejected in the single-star
scenario could still be bound to the binary system, and
therefore continue to evolve chaotically following their
"ejection" in the tidal encounter. In this case, the planet
might survive or suffer a collision just as any other planet
in the system. These "shaved", heated planets might be
much brighter than their cooling age. A direct physi-
cal collision with a MS star, (in our case, the secondary)
should also lead to destruction, as the tidal radius of
Sun-like stars is comparable to their physical radius for
Jupiter-like planets. Physical collisions might also be di-
rectly observable with upcoming transient surveys (Met-
zger et al. 2012).
In the case of an encounter with the primary as a WD,
the tidal radius is much larger than the physical radius of
the WD; thus disruption is likely prior to a direct physi-
cal collision. Disruption can lead to the formation of an
accretion disk around the compact object, and the infall
of material onto it. Such debris disks around WDs po-
tentially serve as a source for long lasting metal pollution
(e.g. Klein et al. 2011 and references therein). Similar
encounters have been studied in the context of the tidal
disruption of a star by a massive black hole (e.g. Rees
1988).
5.3. Collisions with AGB Atmospheres
AGB stars spend a short time (∼ 105 yrs) inflated to
radii of order 1 AU. Consequently we find that the ra-
dius distribution of collisions with the primary is highly
bimodal: they either occur at radii near the maximum,
or the minimum radius in our models -- the tidal radius of
the WD (see Figure 1). Figure 6 shows the distribution
of collision radii and times for all collisions in our fiducial
models. We measure time in units of the host binary's
orbital period at the onset of instability, and t = 0 is set
to the moment when the system crosses the instability
threshold. We can identify several trends. First, the bi-
modality of the distribution is easily observed: only 0.1%
of collisions occur at intermediate radii. Secondly, we can
see that closer binaries are destabilized earlier, and there-
fore collide with the AGB atmosphere at slightly smaller
sizes while the atmosphere is still expanding (note the
blue points are shifted to smaller radii compared to the
green and red points). Lower mass secondaries produce
AGB collisions only at close separations, whereas more
massive secondaries drive the system unstable quickly
enough to drive AGB encounters at all binary separations
considered. We see that most collisions occur within the
first ∼ 200 binary orbits, as seen by Holman & Wiegert
(1999) and David et al. (2003). We discuss collision
timescales in more detail in the Section 5.4.
There are several plausible physical outcomes following
collisions between the planet and AGB star. Collisions
with a partially evolved star might lead to the engulfment
of the planet in the stellar envelope, and ultimately to its
in-spiral and disruption inside the star. In the process,
the planet might transfer energy, angular momentum,
7
and heavy elements into the stellar envelope.
Collisions with the more loosely bound AGB atmo-
sphere likely lead to a qualitatively different outcome.
Such interactions might form a low mass cataclysmic bi-
nary system as the planet penetrates the AGB envelope
before its ejection and accretes from it (Livio & Soker
1984). It can also affect the structure of the planetary
nebulae that ultimately forms around the WD remnant
(Soker 2001), in a similar manner to stellar binary com-
panions. The planet might also completely eject the en-
velope, depending on the its penetration depth, some-
what similar to the case of a common envelope scenario.
In this case, kinetic energy dissipated in the collision and
envelope ejection could be sufficient to capture the sur-
viving planet into a close orbit around the stellar rem-
nant. This latter scenario provides a unique possibility
for placing planets in tight orbits around white dwarfs
(see also Bear et al. 2011 for a related discussion). Note
that if tidal dissipation leads to orbits near the tidal ra-
dius of the WD, such systems would reside in the WD
habitable zone (Agol 2011).
5.4. Varying mass loss rates
Mass loss due to stellar winds occurs on a timescale
that is slow (adiabatic) compared to the planet and stel-
lar orbital period. However, there is another relevant
timescale in the calculation: the average time for a planet
to reach its final "outcome" after becoming unstable, i.e.
collision or a quasi-stable orbit. If the stellar mass and
binary separation evolve on this timescale, then the prob-
ability for a collision or capture is determined by the in-
tegrated probabilities at each state over the time spent
at each state.
In order to isolate the effect of system mass ratio and
separation, and thus determine probabilities as a func-
tion of system parameters exclusively, we run two sets
of comparison simulations with unrealistically long mass
loss timescales set to a constant mass loss rate over 2
Myr, and 20 Myr. We fix the collision radius to the
solar radius at all times.
In both cases, we compute
the timescale over which 75% of particles destined to
collide suffer a collision; this serves as our "equilibrium
timescale," and ranges from 0.2 <∼ tcoll
<∼ 1 Myr, depend-
ing on the system. Changing the fraction of collisions
which defines our equilibrium timescale by 10% has min-
imal effect on the timescales. Thus only the 20 Myr mass
loss timescale case evolves adiabatically with respect to
all timescales in the problem.
In this limit, nearly all (99%) of particles suffer a col-
lision in all of the systems that become unstable. That
collisions increase in the long mass loss timescale limit
is easily seen from the evolution of the planets' CJ . At
the onset of instability, the L1 bottleneck is near its peak
width. This facilitates exchanges between stars, which in
turn are responsible for setting planets on collisional or-
bits. Capture occurs as the bottleneck closes off, which
only happens at the tail end of mass loss. In the case
where the mass loss timescale is long compared to the
collisional time, all of the particles will undergo suffi-
cient exchanges to collide prior to the closing of the zero-
velocity curves, and therefore few are left at the end of
the evolution. We now use these long timescale runs to
show that the relative collision probability between the
8
Fig. 6. -- Distribution of collision times and radii for the fiducial set of runs with realistic radii and mass loss rates. The inset panel is
zoomed in on the radius range of the AGB atmosphere. Masses are in unites of M(cid:12), and separations in AU. Color indicates the initial
binary separation, and symbol size the mass of the secondary. Time is measured in units of each host binary's orbital period, with t = 0
set to be the point at which the system violates the instability criterion in Equation (1). Collisions with the primary are indicated by
circles, while those with the secondary are shown with squares. A movie which shows the population of this phase space is available at
http://www.cfa.harvard.edu/kkratter/BinaryPlanets/radcollmovie.mpg.
primary and secondary scales inversely with mass ratio.
5.5. Collision Rates: Long mass loss times
Once planets become unstable, in the process of bounc-
ing back and forth, they sample the orbital parameter
space about each component, defined by the area en-
closed within the zero velocity curves. We find that the
relative collision rates between the stars are well corre-
lated with the area of the stable region around each star
as defined by the critical radius in Equation (1). When
the stability radius Rs or Rp is smaller, the physical stel-
lar radius subtends a larger fraction of the allowed orbital
area, and thus there is a higher probability for suffering
a collision. The area ratio depends on the square of the
stellar mass ratios. Collision probabilities scale inversely
with the area around the star as shown in Figure 7, mean-
ing that collisions occur predominantly with the star that
is less massive.
In these runs, the stellar radii are fixed at 1R(cid:12)
throughout the calculation. We also run the same models
with the collision radius reduced by a factor of 5, and find
that the collision probabilities remain at 99% across all
unstable systems, although the collision times are shifted
to slightly larger values, as shown in Figure 8. Our inter-
pretation of this trend is that particles set on collisional
trajectories have nearly radial orbits. This property, and
the corresponding high collision rate might well be the
result of considering only coplanar planets.
5.6. Free floating planets: ejection
In the results presented here, destabilized planets ei-
ther collided with one of the binary components, or ex-
changed hosts. No planets were ejected from the system.
Fig. 7. -- Collisional probability for m1 and m2 as a function
of the relative area of the secondary stability region Rs to the to-
tal available stable area: R2
s). These probabilities are
calculated for the simulation set with slow mass loss (20 Myrs),
and constant radius. The lower bound on area for which we show
collisions illustrates the minimum relative areas achieved for com-
parable primary and secondary masses.
p + R2
s/(R2
This is because the minimum CJ reached in the systems
always prohibited escape. For lower CJ , zero-velocity
curves take on a "c" shape, opening up to the circumbi-
nary orbital region, thus allowing escape. In several ex-
ploratory integrations of non-coplanar systems (i ∼ 10◦)
and systems with more unequal initial binary mass ra-
tios (µ ∼ 0.1), we found that planetary ejections are
9
Since the distribution of planet orbital properties is un-
known, we consider three cases : (1) period distributions
similar to that of low mass binaries (log-normal), (2)
period distributions similar to that of high mass bina-
ries (log-uniform), and (3) uniform distribution of peri-
ods. The stability criteria derived by Holman & Wiegert
(1999) does not account for the planetary eccentricity,
and we therefore do not specify it.
For the purposes of calculating frequencies, we inte-
grate our planet distribution down to 3R(cid:12) from the pri-
mary star, however we consider all planets at initial sep-
arations below 7 AU to be "stable", in that they may
interact with the stellar envelope or wind, and thus will
not follow the simple adiabatic mass loss evolution envi-
sioned here (Villaver & Livio 2009).
For all wider systems, we use Equation (1) in Holman
& Wiegert (1999) (which includes the binary eccentric-
ity dependence) to verify whether the initial planetary
system is stable at birth, and if it is not, we generate
a replacement system. We construct 1000 initially sta-
ble planetary systems for each primary mass bin ranging
from 1 − 3M(cid:12) for each of the three planetary orbital
distributions. Each primary mass bin is 0.1 M(cid:12).
We then use the initial to final mass relation deter-
mined by Salaris et al. (2009), to find the final mass of
the primary star as it becomes a WD. Using Equation (2)
to calculate the evolved orbits of the binary and plane-
tary companions, we check whether the planetary system
has crossed the instability threshold.
The total fraction of unstable systems is recorded; the
results are shown in Fig. 9. This figure provides an es-
timate for the fraction of destabilized planetary systems
for a given primary mass host. Only primaries more mas-
sive than 1M(cid:12) become WDs in a Hubble time, and we
therefore use this as a lower limit. Overall we find that
a few up to ∼ 10% of all WD binary systems that con-
tained planets at birth might undergo the TEDI.
Fig. 9. -- The frequency of destabilized planetary systems as a
function of the primary host mass. Three cases are shown, corre-
sponding to a log-normal, log-uniform and uniform distribution of
planetary orbital periods.
7. SUMMARY
In this paper we have examined the effect of binary
stellar evolution on the fate of wide-orbit planets. We
Fig. 8. -- Distribution of collision times for the slow mass loss
(20 Myrs) runs, with collision radius set to R(cid:12) (red), and R(cid:12)/5.
(green). As expected, the collision rate peaks slightly earlier for
larger stellar cross section. Nevertheless, the relative similarity
demonstrates that colliding particles are on sufficiently low angular
momentum orbits that collisions destined to reach a stellar radius
will penetrate to 1/5th of that distance.
possible. For lower mass secondaries, lower CJ may be
reached because the secondary and planet can be more
tightly packed prior to instability. Non-coplanar orbits
likely allow for lower CJ through an expanded phase
space in which eccentricity (and inclination) excitation
increase the velocity of the test particle at a fixed loca-
tion. We defer further discussion of escaping planets to
future work.
6. FREQUENCY OF DESTABILIZED PLANETS
We have thus far shown that the planetary TEDI can
produce a wide range of astrophysically interesting con-
figurations. We now try to estimate the occurrence rate
of planet destabilization. At present, there are few con-
straints on both the overall frequency of planetary sys-
tems in binaries, and on wide orbit systems in general.
Nevertheless, we can calculate the frequency of desta-
bilized systems for various planetary orbital distribu-
tions. Given the simplified approach we take, the results
should be interpreted with caution; they are dependent
on highly uncertain model assumptions.
Following Perets & Kratter (2012) we construct a pop-
ulation of stellar binaries, whose properties are chosen
randomly from the appropriate observed distributions
for A-G stars (Raghavan et al. 2010). The periods are
chosen from a distribution of orbital separations that
is log-normal. The mass ratio of the binary, m1/m2,
is chosen from a Gaussian distribution (mean of 0.6
and dispersion of 0.1). Following Fabrycky & Tremaine
(2007), the distribution of initial eccentricities is a func-
tion of period (Duquennoy & Mayor 1991). For peri-
ods shorter than 1000 days, the eccentricity was chosen
from a Rayleigh distribution [dp ∝ e exp(−βe2)de] with
= β−1/2 = 0.33. For periods longer than 1000
days, the eccentricity was chosen from an Ambartsum-
ian distribution (dp = 2ede), corresponding to a uniform
distribution on the energy surface in phase space.
(cid:10)e2(cid:11)1/2
We populate the binary systems with single planets.
10
have shown that a substantial fraction of binary plane-
tary systems may be susceptible to the planetary version
of the Triple Evolution Dynamical Instability (Perets &
Kratter 2012), where mass loss driven orbital expansion
causes planetary orbits to become unstable and chaotic.
We have considered circular binary systems and coplanar
planets so that we may exploit the simplicity provided by
the circular restricted three-body problem.
As a result of the chaotic orbits induced by mass loss,
planets can suffer close encounters or collide with ei-
ther component in the binary. When close encounters
occur while the mass losing star is on the asymptotic
giant branch, dissipation of energy in the envelope can
potentially lead to planetary captures on close orbits.
Such planets might reside near the tidal radius of the
white dwarf following envelope ejection; these planets
would fall within the white dwarf habitable zone, and
are potentially discoverable with future transit surveys
(Agol 2011). We note that tidal heating might provide
a source of geothermal energy to such planets. Recent
numerical work by Kim & Taam (2012) also finds that
the mere proximity of a planet to the low density atmo-
sphere might excite spiral arm structure observable with
ALMA. Encounters within a few tidal radii of either the
main-sequence secondary, or the newborn white dwarf
likely result in planetary disruption. Planetary disrup-
tion around the white dwarf might contribute to the ob-
served atmospheric pollution (Farihi et al. 2009; Dufour
et al. 2010; Klein et al. 2011; Bochkarev & Rafikov 2011).
Planets that do not suffer close stellar encounters can
become captured by the secondary. Mass loss and planet
hopping drive a fall and then rise in the planets' Jacobi
constants. Mass loss first allows planets to hop back and
forth between the stars; in some cases the late time rise
in the Jacobi constant causes planets to become captured
in orbit about the secondary on quasi-stable orbits (see
Figure 2). Because such planets reside near the boundary
of instability, when the lower mass star in the binary
evolves, a second round of instability is possible.
We suggest that future surveys for planets in evolved
binary systems might find:
• planets around the main-sequence star in a WD-
main-sequence binary at the edge of orbital stabil-
ity
• planets around the WD in a WD-main-sequence
binary on sufficiently close orbits that they would
have been engulfed in AGB atmospheres in stan-
dard evolutionary models
• planets around the more evolved WD in a WD-WD
binary at the edge of orbital stability.
Future work must constrain in detail the outcomes of
collisions between planets and AGB atmospheres to make
more detailed predictions for the types of orbits produced
by this interaction. Similarly, an exploration of multi-
planet, and non-coplanar systems is required to provide
more robust observational predictions.
The authors would like to thank Dimitri Veras and
Andrew Youdin for comments on an earlier draft of this
manuscript, and Nick Moeckel, Matt Holman, Smadar
Naoz, Fred Rasio, and Greg Laughlin for helpful discus-
sions. The authors thank the referee for helpful com-
ments. KMK is supported by an Institute for Theory
and Computation fellowship through the Harvard Col-
lege Observatory. HBP is a BIKURA (FIRST) and CfA
prize fellow.
REFERENCES
Agol, E. 2011, ApJ, 731, L31+
Astakhov, S. A., Burbanks, A. D., Wiggins, S., & Farrelly, D.
2003, Nature, 423, 264
Bear, E., Soker, N., & Harpaz, A. 2011, ApJ, 733, L44
Bochkarev, K. V., & Rafikov, R. R. 2011, ApJ, 741, 36
Bonavita, M., & Desidera, S. 2007, A&A, 468, 721
David, E.-M., Quintana, E. V., Fatuzzo, M., & Adams, F. C.
2003, PASP, 115, 825
Debes, J. H., & Sigurdsson, S. 2002, ApJ, 572, 556
Dufour, P., Kilic, M., Fontaine, G., Bergeron, P., Lachapelle,
F.-R., Kleinman, S. J., & Leggett, S. K. 2010, ApJ, 719, 803
Duquennoy, A., & Mayor, M. 1991, A&A, 248, 485
Eggenberger, A., Udry, S., & Mayor, M. 2004, A&A, 417, 353
Faber, J. A., Rasio, F. A., & Willems, B. 2005, Icarus, 175, 248
Fabian, A. C., Pringle, J. E., & Rees, M. J. 1975, MNRAS, 172,
15P
Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298
Farihi, J., Jura, M., & Zuckerman, B. 2009, ApJ, 694, 805
Fatuzzo, M., Adams, F. C., Gauvin, R., & Proszkow, E. M. 2006,
PASP, 118, 1510
Guillochon, J., Ramirez-Ruiz, E., & Lin, D. 2011, ApJ, 732, 74
Hadjidemetriou, J. D. 1963, Icarus, 2, 440
Heppenheimer, T. A., & Porco, C. 1977, Icarus, 30, 385
Holman, M. J., & Wiegert, P. A. 1999, AJ, 117, 621
Howard, A. W., Marcy, G. W., Bryson, S. T., Jenkins, J. M.,
Rowe, J. F., Batalha, N. M., Borucki, W. J., Koch, D. G.,
Dunham, E. W., Gautier, III, T. N., Van Cleve, J., Cochran,
W. D., Latham, D. W., Lissauer, J. J., Torres, G., Brown,
T. M., Gilliland, R. L., Buchhave, L. A., Caldwell, D. A.,
Christensen-Dalsgaard, J., Ciardi, D., Fressin, F., Haas, M. R.,
Howell, S. B., Kjeldsen, H., Seager, S., Rogers, L., Sasselov,
D. D., Steffen, J. H., Basri, G. S., Charbonneau, D.,
Christiansen, J., Clarke, B., Dupree, A., Fabrycky, D. C.,
Fischer, D. A., Ford, E. B., Fortney, J. J., Tarter, J., Girouard,
F. R., Holman, M. J., Johnson, J. A., Klaus, T. C., Machalek,
P., Moorhead, A. V., Morehead, R. C., Ragozzine, D.,
Tenenbaum, P., Twicken, J. D., Quinn, S. N., Isaacson, H.,
Shporer, A., Lucas, P. W., Walkowicz, L. M., Welsh, W. F.,
Boss, A., Devore, E., Gould, A., Smith, J. C., Morris, R. L.,
Prsa, A., & Morton, T. D. 2011, ArXiv e-prints 1103.2541
Hurley, J. R., Pols, O. R., & Tout, C. A. 2000, MNRAS, 315, 543
Hut, P., Makino, J., & McMillan, S. 1995, ApJ, 443, L93
Johnson, J. A., Fischer, D. A., Marcy, G. W., Wright, J. T.,
Driscoll, P., Butler, R. P., Hekker, S., Reffert, S., & Vogt, S. S.
2007, ApJ, 665, 785
Kim, H., & Taam, R. E. 2012, ApJ, 744, 136
Klein, B., Jura, M., Koester, D., & Zuckerman, B. 2011, ApJ,
741, 64
Livio, M., & Soker, N. 1984, MNRAS, 208, 763
Luk'yanov, L. G. 2009, Astronomy Letters, 35, 349
Metzger, B. D., Giannios, D., & Spiegel, D. S. 2012, ArXiv
e-prints 1204.0796
Moeckel, N., & Veras, D. 2012, MNRAS, 2631
Mudryk, L. R., & Wu, Y. 2006, ApJ, 639, 423
Mugrauer, M., & Neuhauser, R. 2009, A&A, 494, 373
11
Murray, C. D., & Dermott, S. F. 2000, Solar System Dynamics,
Salaris, M., Serenelli, A., Weiss, A., & Miller Bertolami, M. 2009,
ed. Murray, C. D. & Dermott, S. F.
Nordhaus, J., Spiegel, D. S., Ibgui, L., Goodman, J., & Burrows,
A. 2010, MNRAS, 408, 631
Perets, H. B. 2010, ArXiv e-prints 1001.0581
Perets, H. B., & Kratter, K. M. 2012, ArXiv e-prints 1203.2914
Rabl, G., & Dvorak, R. 1988, A&A, 191, 385
Raghavan, D., McAlister, H. A., Henry, T. J., Latham, D. W.,
Marcy, G. W., Mason, B. D., Gies, D. R., White, R. J., & ten
Brummelaar, T. A. 2010, ApJS, 190, 1
Rees, M. J. 1988, Nature, 333, 523
Ruskol, E. L. 1961, Soviet Ast., 4, 657
ApJ, 692, 1013
Soker, N. 2001, MNRAS, 324, 699
Veras, D., & Tout, C. A. 2012, MNRAS, 2678
Veras, D., Wyatt, M. C., Mustill, A. J., Bonsor, A., & Eldridge,
J. J. 2011, MNRAS, 417, 2104
Villaver, E., & Livio, M. 2007, ApJ, 661, 1192
-- . 2009, ApJ, 705, L81
Wisdom, J. 1980, AJ, 85, 1122
We include in tabular form the outcomes for all binaries in our fiducial models. We classify particles according to
three possible outcomes: bound, colliding, and tidally interacting with one of the two components. Note that in several
cases, the total number of interactions sums to greater than the number of particles in the simulation (100). This is
due to the fact that several planets suffer tidal encounters with both components. We discuss these results in Section
5.1.
APPENDIX
12
Tabulated results from the fiducial suite of integrations with 100 test
particles in each binary. Particles are classified as either bound,
colliding, or tidally interacting with either the primary or secondary.
TABLE 1
Mass (M(cid:12))
Sep. (AU) Bound 1 Bound 2 Coll 1 Coll 2 Tid 1 Tid 2
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.6
0.6
0.6
0.6
0.6
0.6
0.6
0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.8
0.8
0.8
0.8
0.8
0.8
0.8
0.9
0.9
0.9
0.9
0.9
0.9
0.9
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.3
1.3
1.3
1.3
1.3
1.3
1.3
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.7
1.7
1.7
1.7
1.7
1.7
1.7
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
75
80
85
90
95
100
105
0.0
100.0
100.0
100.0
100.0
100.0
100.0
0.0
0.0
0.0
100.0
100.0
100.0
100.0
0.0
0.0
0.0
0.0
0.0
100.0
100.0
0.0
0.0
0.0
1.0
0.0
0.0
50.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
3.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
2.0
0.0
0.0
0.0
0.0
0.0
2.0
9.0
6.0
2.0
0.0
0.0
0.0
1.0
0.0
14.0
10.0
14.0
3.0
0.0
1.0
6.0
3.0
18.0
20.0
15.0
10.0
0.0
0.0
4.0
12.0
14.0
23.0
18.0
0.0
2.0
5.0
4.0
13.0
18.0
26.0
0.0
0.0
1.0
7.0
16.0
15.0
25.0
0.0
1.0
0.0
1.0
6.0
17.0
14.0
34.0
0.0
0.0
0.0
0.0
0.0
0.0
34.0
29.0
29.0
0.0
0.0
0.0
0.0
57.0
33.0
37.0
44.0
40.0
0.0
0.0
66.0
51.0
50.0
41.0
44.0
69.0
33.0
69.0
77.0
56.0
63.0
48.0
65.0
67.0
71.0
71.0
75.0
68.0
62.0
64.0
59.0
74.0
85.0
80.0
81.0
73.0
58.0
62.0
77.0
82.0
83.0
87.0
78.0
74.0
66.0
80.0
91.0
92.0
88.0
71.0
83.0
73.0
89.0
93.0
96.0
92.0
89.0
77.0
83.0
50.0
0.0
0.0
0.0
0.0
0.0
0.0
51.0
50.0
57.0
0.0
0.0
0.0
0.0
32.0
48.0
46.0
37.0
43.0
0.0
0.0
29.0
38.0
36.0
42.0
40.0
17.0
9.0
23.0
17.0
26.0
23.0
32.0
24.0
17.0
22.0
21.0
20.0
14.0
11.0
17.0
21.0
23.0
15.0
16.0
7.0
11.0
17.0
18.0
20.0
13.0
10.0
9.0
9.0
7.0
7.0
17.0
5.0
7.0
5.0
13.0
2.0
2.0
6.0
5.0
4.0
7.0
5.0
6.0
3.0
11.0
0.0
0.0
0.0
0.0
0.0
0.0
8.0
8.0
10.0
0.0
0.0
0.0
0.0
2.0
10.0
11.0
12.0
5.0
0.0
0.0
0.0
0.0
2.0
7.0
7.0
7.0
4.0
0.0
2.0
1.0
1.0
2.0
5.0
7.0
0.0
0.0
0.0
0.0
2.0
3.0
3.0
0.0
0.0
0.0
0.0
1.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
5.0
0.0
0.0
0.0
0.0
0.0
0.0
7.0
13.0
4.0
0.0
0.0
0.0
0.0
8.0
8.0
6.0
7.0
12.0
0.0
0.0
4.0
4.0
6.0
7.0
9.0
7.0
4.0
8.0
4.0
6.0
4.0
3.0
3.0
9.0
6.0
3.0
3.0
0.0
5.0
1.0
5.0
3.0
0.0
2.0
0.0
1.0
4.0
1.0
3.0
3.0
2.0
0.0
0.0
0.0
2.0
3.0
4.0
0.0
1.0
1.0
0.0
1.0
5.0
1.0
0.0
0.0
1.0
1.0
0.0
|
1903.06130 | 1 | 1903 | 2019-03-14T17:03:37 | Hot exozodiacal dust: an exocometary origin? | [
"astro-ph.EP"
] | We aim to explore two exozodiacal dust production mechanisms, first re-investigating the Poynting-Robertson drag pile-up scenario, and then elaborating on the less explored, but promising exocometary dust delivery scenario. We developped a new versatile, numerical model that calculates the dust dynamics, with non orbit-averaged equations for the grains close to the star. The model includes dust sublimation and incorporates a radiative transfer code for direct comparison to the observations. We consider in this study four stellar types, three dust compositions, and we assume a parent belt at 50 au. We find that, in the case of the Poynting-Robertson drag pile-up scenario, it is impossible to produce long-lived submicron-sized grains close to the star. The inward drifting grains fill in the region between the parent belt and the sublimation distance, producing an unrealistically strong mid-infrared excess compared to the near-infrared excess. The dust pile-up at the sublimation radius is by far insufficient to boost the near-IR flux of the exozodi to the point where it dominates over the mid-infrared excess. In the case of the exocometary dust delivery scenario, we find that a narrow ring can form close to the sublimation zone, populated with large grains several tens to several hundred of micrometers in radius. Although not perfect, this scenario provides a better match to the observations, especially if the grains are carbon-rich. We also find that the required number of active exocomets to sustain the observed dust level is reasonable. We conclude that the hot exozodiacal dust detected by near-infrared interferometry is unlikely to result from inwards grains migration by Poynting-Robertson drag from a distant parent belt, but could instead have an exocometary origin. [Abridged] | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. V_6.1
March 15, 2019
c(cid:13)ESO 2019
9
1
0
2
r
a
M
4
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
3
1
6
0
.
3
0
9
1
:
v
i
X
r
a
Hot exozodiacal dust : an exocometary origin?
É. Sezestre1, J.-C. Augereau1, and P. Thébault2
1 Univ. Grenoble Alpes, CNRS, IPAG, 38000 Grenoble, France
2 LESIA, Observatoire de Paris, CNRS, Université Paris Diderot, Université Pierre et Marie Curie, 5 place Jules Janssen, 92190
Meudon, France
Received -; accepted -
ABSTRACT
Context. Near- and mid-infrared interferometric observations have revealed populations of hot and warm dust grains populating the
inner regions of extrasolar planetary systems. These are known as exozodiacal dust clouds, or exozodis, reflecting the similarity with
the Solar System's zodiacal cloud. Radiative transfer models have constrained the dust to be dominated by tiny submicron-sized,
carbon-rich grains that are accumulated very close to the sublimation radius. The origin of this dust is an unsolved issue.
Aims. We aim to explore two exozodiacal dust production mechanisms, first re-investigating the Poynting-Robertson drag pile-up
scenario, and then elaborating on the less explored, but promising exocometary dust delivery scenario.
Methods. We developped a new versatile, numerical model that calculates the dust dynamics, with non orbit-averaged equations for
the grains close to the star. The model includes dust sublimation and incorporates a radiative transfer code for direct comparison to
the observations. We consider in this study four stellar types, three dust compositions, and we assume a parent belt at 50 au.
Results. We find that, in the case of the Poynting-Robertson drag pile-up scenario, it is impossible to produce long-lived submicron-
sized grains close to the star. The inward drifting grains fill in the region between the parent belt and the sublimation distance,
producing an unrealistically strong mid-infrared excess compared to the near-infrared excess. The dust pile-up at the sublimation
radius is by far insufficient to boost the near-IR flux of the exozodi to the point where it dominates over the mid-infrared excess. In
the case of the exocometary dust delivery scenario, we find that a narrow ring can form close to the sublimation zone, populated with
large grains several tens to several hundred of micrometers in radius. Although not perfect, this scenario provides a better match to
the observations, especially if the grains are carbon-rich. We also find that the required number of active exocomets to sustain the
observed dust level is reasonable.
Conclusions. We conclude that the hot exozodiacal dust detected by near-infrared interferometry is unlikely to result from inwards
grains migration by Poynting-Robertson drag from a distant parent belt, but could instead have an exocometary origin.
Key words. Planetary systems -- Circumstellar matter -- Methods: numerical -- Infrared: planetary systems -- Comets: general --
Zodiacal dust
1. Introduction
Hot exozodiacal dust ("exozodis") has been detected, by means
of interferometric observations in the near-infrared (near-IR, H-
or K-band), around about 25 main sequence stars (Absil et al.
2013; Ertel et al. 2014, 2016; Kral et al. 2017; Nuñez et al. 2017).
These exozodis are very bright, amounting to ∼ 1% of the stel-
lar flux in the K-band, which is about 1000 times more than the
solar system's own zodiacal cloud in the same spectral range.
For some of these systems, a "warm" counterpart has also been
detected in the mid-infrared (mid-IR, 8-20 µm, e.g. Mennesson
et al. 2013; Su et al. 2013; Mennesson et al. 2014; Ertel et al.
2018), but this mid-IR exozodi to star flux ratio never exceeds
the flux ratios in the H- or K-band (Kirchschlager et al. 2017).
Furthermore, for the handful of systems for which parametric
modelling based on radiative transfer codes has been performed
(Absil et al. 2006; Di Folco et al. 2007; Absil et al. 2008; Ake-
son et al. 2009; Defrère et al. 2011; Lebreton et al. 2013; Kirch-
schlager et al. 2017), the ratio between the fluxes in the near-IR
and mid-IR has constrained the dust to be dominated by tiny
submicron-sized grains that are accumulated very close to the
sublimation radius (hereafter rs, typically a few stellar radii).
Send offprint requests to: É. Sezestre,
email: [email protected]
The presence of such large amounts of very small grains so
close to their star poses a challenge when it comes to explaining
the exozodis' origin. Indeed, the canonical explanation invoked
for "standard" cold debris disks, i.e., the in situ steady produc-
tion of small grains by a collisional cascade starting from larger
parent bodies (e.g. Krivov 2010), cannot hold here because col-
lisional erosion is much too fast in these innermost regions to
be sustained over periods comparable to the system's age (Bon-
sor et al. 2012; Kral et al. 2017). Therefore, the long-term ex-
istence of a hot exozodi requires both an external reservoir of
material and an inward transport mechanism, feeding with dust
the region close to the sublimation radius at a rate of about
10−10 − 10−9 M⊕/year (e.g. Absil et al. 2006; Kral et al. 2017).
A significant fraction (more than ∼20%) of nearby solar- and A-
type stars does possess an extrasolar analog to the Kuiper belt
(Montesinos et al. 2016; Sibthorpe et al. 2018; Thureau et al.
2014), indicating that external reservoirs for exozodis are com-
mon. The inward transport mechanism must then be sufficiently
generic to affect more than 10% of the nearby stars, indepen-
dent of their age and spectral type (Ertel et al. 2014; Nuñez et al.
2017). For instance, large-scale dynamical instabilities in plan-
etary systems, that could occur randomly (e.g. the Late Heavy
Bombardment in the Solar System), were shown to indeed sig-
nificantly increase the number of small bodies scattered from an
Article number, page 1 of 21
A&A proofs: manuscript no. V_6.1
external Kuiper-like belt toward the star, but because each event
lasts less than a few million years, the probability to observe
hot exozodiacal dust produced during such an event is less than
0.1% (Bonsor et al. 2013). This mechanism cannot explain the
vast majority of the hot exozodis.
So far, two main categories of exozodi-origin scenarios have
been explored. The first one assumes that the dust is collisionally
produced further out in the system (in an asteroidal or Kuiper-
like belt) and migrates inward, because of Poynting-Robertson
drag (hereafter PR-drag), until it reaches the sublimation dis-
tance rs. There, it starts to sublimate and shrink until radia-
tion pressure becomes significant and increases its orbital semi-
major axis and eccentricity, while keeping its periastron nearly
the same. This will slow-down the inward migration and thus po-
tentially create a pile-up of small grains close to rs. This scenario
follows the pionieering work of Belton (1966) predicting a den-
sity peak near the sublimation distance in the Solar System, and
those by Mukai et al. (1974) and Mukai & Yamamoto (1979)
attempting to explain the observed flux bump at about 4 R(cid:12) in
the F-corona (the hot component of the zodiacal dust cloud).
However, the estimated amplitude of this pile-up seems to be
too weak to explain the observed near-IR excesses in extrasolar
systems (Kobayashi et al. 2008, 2009, 2011; Van Lieshout et al.
2014). Another problem is that this scenario does not seem to be
able to produce grains that are as small as those derived from ra-
diative transfer modeling. However, it is worth noting that these
results were obtained using orbit-averaged equations of motion
that might become inaccurate close to rs because of the very fast
variations imposed by the sublimation.
A second way of delivering dust in the innermost regions
of planetary systems is by the sublimation of large asteroidal or
cometary bodies, originating in an external belt, and scattered
inwards by a chain of low-mass planets (Bonsor et al. 2012,
2014; Raymond & Bonsor 2014; Marboeuf et al. 2016). There
are evidences for exocometary activity around other stars than
the Sun, through the observation of transient, Doppler-shifted
gas absorption lines (e.g. Beust & Morbidelli 2000; Kiefer et al.
2014a,b, and references therein), and the analysis of Kepler tran-
sit light curves attributed to trailing dust tails passing in front of
the star (Kiefer et al. 2017; Rappaport et al. 2018, with mass loss
rates of ∼ 10−12 M⊕/year and > 10−10 M⊕/year, respectively).
In the Solar System, comets are supposed to contribute signifi-
cantly to the zodiacal cloud (e.g. Liou et al. 1995; Dermott et al.
1996). Nesvorný et al. (2010) estimated for example that ∼ 90%
of the zodiacal dust originates from Jupiter family comets. The
cometary hypothesis as a source of hot exozodiacal dust has,
however, never been tested quantitatively in terms of the level
of dustiness that can be obtained near the rs region.
This paper reinvestigates both these scenarios. For the PR-
drag case (Sec. 3), we use for the first time a sophisticated nu-
merical model that does not rely on orbit-averaged equations
in the crucial sublimation region (Sec. 2). We also explore the
potential role played by the Differential Doppler Effect (DDE)
evoked by Kimura et al. (2017, 10th meeting on Cosmic Dust,
Tokyo)1. As for the comet-delivery case, we perform the first
quantitative exploration of this scenario in the context of exo-
zodis, following the fate of the dust that is produced as the comet
sublimates (Sec. 4). For each scenario, we explore a wide range
of possible grain compositions and stellar types (Sec. 2). Rather
than checking the validity of each scenario by assessing how
well they can reproduce the predictions of radiative-model fits
(grain location and typical sizes), we chose to directly focus on
1 https://www.cps-jp.org/~dust
Article number, page 2 of 21
the observational constraints themselves, in particular the fluxes
in the near- and mid-IR.
2. Numerical Model
2.1. General philosophy
We use in essence the same numerical code to investigate both
the PR-drag pile-up and the cometary delivery scenarios. Our
model performs a consistent treatment of a grain evolution, from
its release to its ejection, sublimation, or fall onto the star. We
take into account stellar gravity, stellar radiation/wind pressure,
PR-drag and sublimation. In a more advanced version, the stellar
magnetic field can be turned on, but this capability will not be
used in this paper.
In this study, we chose to neglect collective effects such as
mutual collisions. This might appear as a step back when com-
pared to the studies of Kobayashi et al. (2009) and Van Lieshout
et al. (2014), who did take into account collisional effects (albeit
in a very simplified way) for the PR-drag pile-up scenario. How-
ever, we believe that this neglect of collisions does not radically
bias our results. Van Lieshout et al. (2014) has indeed shown
that, because of the self-regulating interplay between collisions
and PR-drag, collisional effects will only play a significant role,
potentially halting the inward drift of grains, very close to the lo-
cation of the parent body belt releasing the dust grains. As soon
as the dust has migrated away from the parent belt, its number
density is always low enough for mutual collisions not to have
a major effect on its evolution (see Fig. 2 of Van Lieshout et al.
2014). In this respect, the only drawback of not taking into ac-
count collisions is that we cannot derive the density and the mass
of the dust-producing parent belt, but this is not the main focus of
our study, which concentrates on the evolution of the dust once
it has reached the inner regions of the system.
For the dust evolution in these innermost regions, our code
presents a step forward as compared to previous studies because
it does not rely on orbit-averaged da/dt and de/dt estimates but
integrates the exact equations of motion up until the grain is re-
moved. This is a crucial point in the critical region close to the
sublimation radius, where a grain radius can vary on timescales
much smaller than the local orbital period, thus inducing dynam-
ical changes that cannot be accounted for with averaged esti-
mates. In addition, the orbit-averaged de/dt estimates can lead to
eccentricity values that can be infinitely small, whereas in real-
ity there is always a minimum "residual" osculating eccentricity
below which the particle's orbit cannot go (see section 3.1).
In addition to the dust evolution, the output of the code is a
global density map assuming the system is at steady state. This
is used to produce a synthetic spectrum of the exozodi that can
be compared to mesured spectra, and also flux levels mesured by
interferometric studies. Since the mass of dust close to the star in
the PR-drag scenario is not constrained by the mass of the dust
producing belt, we try to reproduce the trend of the spectra, and
use the mass of the exozodi as a free parameter. More specifi-
cally, we scale the mass such that the excess corresponds to ob-
servations at 2 µm, and we use the excess observed in mid-IR to
discuss the relevancy of the examined scenarios (around 1% at
8 -- 20 µm, e.g. Kirchschlager et al. 2017). On the contrary, in the
cometary release scenario, the flux level can be estimated by the
mass of the releasing comet, giving constraints on its radius.
É. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(a) Carbon
(b) Astrosilicate
(c) Glassy silicate
Fig. 1: βpr for 3 different grain compositions around different spectral type stars (ticks correspond to the grain sizes used in our
simulations). The solid horizontal line is the limit β = 0.5: grains over this value are blown-out by radiation pressure if produced
from circular orbits. The dashed horizontal line is the limit β = 1: grains above this value are always expelled, regardless the of the
way they are produced.
2.2. Dynamical approach
The code computes the dynamics of a set of compact dust
grains with initial sizes chosen to sample different dynamical
behaviours. The equation of motion is solved with a 4th order
Runge-Kutta integrator with an adaptive timestep. The code is
able to take into account the stellar gravity (Fgrav), the radiation
pressure and the Poynting-Roberston drag (FPR), and the Dif-
ferential Doppler Effect (FDDE, e.g. Burns et al. 1979). Each of
these effects can be individually switched on or off at any time.
The forces are expressed as follows :
Fgrav = −GM(cid:63)m
r2
GM(cid:63)m
FPR = βpr
r2
FDDE = − ω(cid:63)R2
4
(cid:63)
(cid:19)
(cid:114)
er − v
1 − r
c
c
GM(cid:63)
(cid:20)(cid:18)
· er
βpr(cid:112)1 − βpr
r5
(cid:21)
· v
c
(1)
(2)
(3)
where er is the radial unit vector, G the gravitational constant, c
the speed of light, M(cid:63) the mass of the star, R(cid:63) the stellar radius,
ω(cid:63) the rotation frequency of the star, m the mass of the grain, r
the distance of the grain to the star, v the grain velocity and r the
radial velocity, βpr the ratio of the radiation pressure force to the
gravitational force.
Other forces, in particular the stellar wind pressure and the
Lorentz force, are implemented in the code but will not be used
in this study. The pressure due to the stellar wind is comparable
to the radiation pressure only for submicron-sized grains around
late-type stars. As we choose to focus on K-type and earlier stars
(Sec. 2.3), we will not take into account the stellar wind pressure.
For consistency and simplicity, we will rename the βpr parameter
as β in the following. The Lorentz force acting on charged grains
interacting with the large-scale stellar magnetic field can also
affect the grain dynamics, as evidenced by Czechowski & Mann
(2012) or Rieke et al. (2016). We will neither discuss the Lorentz
force as it exceeds the scope of this study.
The initial conditions of the simulations depend on the sce-
nario that is considered. These are detailled in Secs. 3 and 4 for
the PR-drag pile-up and the cometary delivery scenarios, respec-
tively. The grain dynamics is computed until one of following
criteria is met:
-- the grain sublimates completely. This is reached when the
grain size is below the lower limit of the predefined size grid,
which in most cases corresponds to a size smaller than 1 nm.
-- the grain falls onto the star. This is assumed to happen when
the distance of the grain to the surface of the star is less than
0.1 R(cid:63).
-- the grain is expelled. This is assumed to occur when the dis-
tance to the star is over 1000 au.
-- the grain is too old. This is considered to be the case when
the integration time is over one million years, meaning the
grain has not evolved.
The integration timestep is taken as a fraction of the local revo-
lution period (typically a hundredth), to ensure a sufficient reso-
lution at every distance from the star. As a test of the code, we
reproduced the results in Figure 5 of Krivov et al. (1998) with
great fidelity, as shown in Appendix A. However, for the grains
released from parent bodies in a distant belt and then migrating
inward by PR-drag, like in the scenario developed in Sec. 3, this
short timestep becomes a numerical limitation. Therefore, and as
long as the grain remains far from the star, we opt in this case for
the orbit-averaged prescription of Wyatt & Whipple (1950, their
Eq. 9) to evolve the grains by PR-drag. This approach allows to
save computational time during the less critical evolution stages
(stage I as defined in Sec. 3.1.1), and is similar to the method-
ology employed by Kobayashi et al. (2009) and Van Lieshout
et al. (2014). According to Wyatt & Whipple (1950), the quan-
tity ae−4/5(1−e2) remains constant during the PR-drag migration,
where a is the grain semi-major axis and e its eccentricity. We
use this conservation principle to estimate the semi-major axis
and the eccentricity as the grain migrates inward until the full,
non orbit-averaged simulation is switched on, in contrast with
what was done in previous studies. The switch is done when
the grain reaches an equilibrium temperature at periastron that
is half its sublimation temperature, to prevent sublimation to oc-
cur during the orbit-averaged phase. We also continuously mon-
itor the evolution of the grain radius due to sublimation during
this phase, in order to stop the orbit-averaged treatment if the ra-
dius is decreased by more than 1% of its initial value. We have
checked on a test run that this approach provides the same results
as those obtained with the full-simulation. It should be noted
that, in the cometary scenario developed in Sec. 4, the whole
grain evolution is done using non orbit-averaged equations.
2.3. Stellar and grain properties
In this paper, we consider four different stellar types, ranging
from A0 to K0. For that purpose, we choose four representa-
Article number, page 3 of 21
10-910-810-710-610-510-410-310-2Grain size (m)10-510-410-310-210-1100101102103βA0F0G0K010-910-810-710-610-510-410-310-2Grain size (m)10-510-410-310-210-1100101102βA0F0G0K010-810-710-610-510-410-310-2Grain size (m)10-510-410-310-210-1100101102βA0F0G0K0A&A proofs: manuscript no. V_6.1
Table 1: Reference stars used in the code. Luminosity and mass
are estimated by the code by interpolating the values computed
for spectral type.
Spec.
type
A0
F0
G0
K0
(pc)
Name
Distance
7.68±0.02a
Vega
ρ Gem 18.05±0.08b
11.82±0.04b
Iam Ser
11.14±0.01b
54 Psc
V-band Luminosity Mass
(M(cid:12))
mag.
0.03c
2.9
4.18c
1.6
4.42d
1.05
5.88d
0.79
(L(cid:12))
57
5.8
2.0
0.57
References. (a) van Leeuwen (2007) (b) Gaia Collaboration et al. (2016,
2018) (c) Ducati (2002) (d) van Belle & von Braun (2009)
used here is similar to the one in Kobayashi et al. (2011), with the
transformations from one set of thermodynamical parameters to
another given in Appendix B. At each timestep in the dynamical
code, the mass lost by a grain due to sublimation is computed,
and the grain size as well as the β value are modified accord-
ingly for the next dynamical timestep. It is worth noting that
the sublimation timescales can be very sensitive to the compo-
sition. In particular, the behaviour of the glassy silicates is very
different from that of the carbon and astrosilicate grains. For ex-
ample, while it takes 2 × 106 and 3 × 106 s to entirely sublimate
a carbon and an astrosilicate grain of 1 µm, respectively, once
the sublimation temperature is reached, a glassy silicate grain of
the same size will sublimate in only 102 s at its own sublimation
temperature.
2.4. Synthetic Spectral Energy Distributions
By combining the different, single-size (single-β) grain runs, we
can estimate a density profile, as parameterized by the vertical
optical depth τ, assuming that the grains are produced at steady
state from the parent belt. The usual method consists in record-
ing the grains positions at regularly spaced time intervals, and
pile-up these different positions following a procedure similar to
Thébault et al. (2012) until the grain is removed from the sys-
tem (ejection, sublimation or fall onto the star). Here, we em-
ploy a different approach to compute τ, described in details in
Appendix C. It combines density profiles derived from the lim-
ited number of test grains for which the dynamics has been cal-
culated accurately, and timescale estimates for a broader range
of grains sizes, to produce 2-D (r, s) density and optical depth
maps. These maps are obtained assuming an initial differential
size distribution proportional to s−3.5.
We also developped a Python implemented version of the
GRaTeR radiative transfer code (Augereau et al. 1999) that al-
lows to calculate thermal emission and scattered light maps at
any wavelength from the 2-D (r, s) maps, as well as spectral
energy distributions (SED) of the exozodis in order to directly
compare our numerical results with the observations.
3. PR-drag pile-up scenario
We consider a set-up similar to the one explored by Kobayashi
et al. (2011) and Van Lieshout et al. (2014), with a population of
small grains assumed to be released by collisions in a Kuiper-belt
like ring (parent bodies located at r0 = 50 au), whose evolution
is then followed, taking into account PR-drag and sublimation
near the star, until the grains leave the system either by total
sublimation, fall onto the star or dynamical ejection. We explore
4 stellar types and 3 different grain compositions (see Tables 1
and 2). We consider 24 initial grain sizes, ranging from 1.7 nm
to 1 mm, and thus 24 different initial β values (vertical tick marks
in Figs. 1a, 1b and 1c). We consider that the grains are released
from parent bodies on circular orbits at r0 = 50 au, so that the
grains' initial orbit is given by a = r0 × (1 − β)/(1 − 2β) and
e = β/(1 − β).
3.1. Grain evolution
3.1.1. General behaviour
Figs. 2 and 3 present the evolution of grain sizes and orbital ele-
ments for a subset of the explored parameter space (stellar type,
grain composition).
tive nearby stars, that do not necessarily possess hot exozodiacal
dust. Their properties are summarized in Tab. 1.
We consider three different grain compositions, parameter-
ized by their physical, optical and thermodynamical properties.
In the following, "carbon" will refer to amorphous carbonaceous
grains, "astrosilicates" and "glassy silicate" to amorphous sili-
cate grains. The latter two compositions differ by their optical
indexes. Optical indexes for carbon grains are taken from Zubko
et al. (1996, ACAR sample), while those for astrosilicates are
from Draine (2003). The optical indexes for glassy silicates com-
bine measurements for obsidian from Lake Co. Oregon (Pollack
et al. 1973; Lamy 1978) in the spectral range 0.1 µm -- 50 µm, with
a constant value for the real part beyond λ=50 µm, and a constant
value from λ=50 µm to 300 µm for the imaginary part, followed
by the imaginary part of the astrosilicates of Draine (2003) be-
yond λ=300 µm. Below λ=0.1 µm, both the real and imaginary
parts are assumed to be constant. This set of optical indexes for
the "glassy silicates" corresponds to the one used in Kimura et al.
(1997) and Krivov et al. (1998) for silicate grains, with the only
addition of the extension beyond λ=300 µm which is specific to
this study.
We employ the Mie theory, valid for hard spheres, to com-
pute the dust optical properties. These are used to derive the β
ratios (e.g. Eq. 3 in Sezestre et al. 2017) and the radial profiles of
the grain temperature (e.g. Eq. 4 in Lebreton et al. 2013). Both
depend on the grain size, on the grain composition and on the
star that is considered, as shown in Figs. 1a, 1b and 1c in the
case of the β ratios.
The sublimation prescription is taken from Lebreton et al.
(2013, their Eqs. 17 and 18), and follows the methodology de-
scribed in Lamy (1974). The evolution of the grain size s reads:
ds
dt
= − α
ρ
kBT
2πµmu
ρeq,
(4)
where ρ is the grain density, kB is the Boltzmann constant, T
is the grain temperature, µ is the mean molecular mass of the
considered dust composition, and mu is the atomic mass unit.
The equilibrium gas density ρeq around the grain is given by:
log10 ρeq = B − A
Tsub
− log10 Tsub,
(5)
with Tsub being the sublimation temperature of the grain. We
have assumed that the pressure of the gas surrounding the grain
(ρgas in Lebreton et al. 2013) is negligible, and the efficiency
factor α to be 0.7 like in Lamy (1974). The thermodynamical
properties are documented in Tab. 2 for each of the three com-
positions considered in this paper. The sublimation prescription
Article number, page 4 of 21
(cid:115)
É. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(a)
(c)
(e)
(b)
(d)
(f)
Fig. 2: PR-drag pile-up scenario: Grain evolution, as a function of initial grain size, for three different stellar types and 2 grain
compositions. Left panels: Grain size as a function of stellar distance. The arrows denote the temporal evolution. The dash-dotted
line is the sublimation distance as a function of grain size, while the vertical dashed line indicates the position of the parent bodies.
The gray horizontal zone identifies the range of grains with β > 1, and the horizontal dotted lines correspond to β = 0.5 (see also
Fig. 1). Right panels: Eccentricity as a function of semi-major axis. The horizontal plain line represents the e=1 limit beyond which
particles are on unbound orbits. For both the left and right panels, the left yellow area corresponds to the physical location of the
star and the arrows denote the evolution way.
Article number, page 5 of 21
10-310-210-1100101102r (au)10-910-810-710-610-510-410-3s (m)A0 star, Astrosilicate10-310-210-1100101102semi-major axis (au)10-610-510-410-310-210-1100eA0 star, Astrosilicate2e-05 m5e-05 m2e-04 m10-310-210-1100101102r (au)10-910-810-710-610-510-410-3s (m)F0 star, CarbonStage IStage IIStage III10-310-210-1100101102semi-major axis (au)10-610-510-410-310-210-1100eF0 star, CarbonStage IStage I bStage IIStage III5e-06 m2e-05 m5e-05 m10-310-210-1100101102r (au)10-910-810-710-610-510-410-3s (m)G0 star, Astrosilicate10-310-210-1100101102semi-major axis (au)10-610-510-410-310-210-1100eG0 star, Astrosilicate2e-09 m5e-09 m2e-08 m2e-06 m5e-06 m2e-05 mTable 2: Grain parameters used in the code. Sublimation parameters A and B refer to those used in Lebreton et al. (2013). Appendix B
provides a comparison with other sublimation formulae and notations used in the literature.
A&A proofs: manuscript no. V_6.1
Name
Density
Mean molecular mass
Sublimation temperature
A
B
Symbol
ρ (kg.m−3)
µ (g.mol−1)
Tsub (K)
(cgs)
(cgs)
Carbon
1.78×103
12.01
2000
37215
7.2294
Astrosilicate Glassy silicate Reference
3.5×103
172.2
1200
28030
12.471
2.37×103
67.00
1200
24918
7.9356
1, 1, 2
1, 1, 2
1
3, 4, 5
3, 4, 5
References. (1) Carbonaceous material and Silicates of Lebreton et al. (2013); (2) Obsidian of Lamy (1974) ; (3) C1 specy of Zavitsanos & Carlson
(1973); (4) Astronomical silicate of Kama et al. (2009); (5) Silicate of Kimura et al. (1997)
(a)
(c)
(b)
(d)
Fig. 3: Same as Fig. 2 but for the G0 and K0 stars, and glassy silicates.
As can be seen, the initial stage (labelled "stage I", after
Kobayashi et al. 2009, and reproduced in Fig. 2c and 2d) is rel-
atively similar for all cases and corresponds to the behaviour
found by previous studies using orbit-averaged equations: the
grain drifts inward due to PR-drag and its orbit is progressively
circularized, while its size remains constant because it is too far
from the sublimation region. We note, however, that, contrary
to the predictions of orbit-averaged prescriptions, the eccentric-
ity stops to decrease at a given point and starts to slowly increase
again as its semi-major axis continues to drop (named Stage Ib in
Fig. 2d). This inflection point corresponds to a "residual" value
Article number, page 6 of 21
below which the osculating eccentricity of the PR-drag drifting
particle cannot fall, which is due to the intrinsic curvature of the
tightly wound spirals that the grain actually follows as it migrates
inward. The osculating eccentricity corresponding to these spi-
rals can be approximated, to a first order, by (da/a)orb, which is
the relative variation of the particle's semi-major axis, due to PR-
drag, over one orbital period as given by the averaged equations
used by Kobayashi et al. (2009) or Van Lieshout et al. (2014).
Taking the right-hand term of Eq. 1 of Kobayashi et al. (2011)
10-310-210-1100101102r (au)10-810-710-610-510-410-3s (m)G0 star, Glassy Silicate10-310-210-1100101102semi-major axis (au)10-610-510-410-310-210-1100eG0 star, Glassy Silicate2e-08 m5e-08 m2e-06 m5e-06 m10-310-210-1100101102r (au)10-810-710-610-510-410-3s (m)K0 star, Glassy Silicate10-310-210-1100101102semi-major axis (au)10-610-510-410-310-210-1100eK0 star, Glassy Silicate5e-08 m2e-07 m5e-07 m2e-06 mÉ. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(drift rate due to PR-drag), we get
da =
2βGM∗
ac
T
(6)
where T is the orbital period and c the speed of light. This leads
to a residual eccentricity of the order of
(cid:114)
eres ≈ da
a
=
4πβ
c
GM∗
a ,
(7)
which indeed increases with decreasing a. This non-zero eres is
always relatively small, less than a few 10−3, but it cannot be
ignored because even such a small value can make a difference
as to the final fate of a grain as it starts sublimating.
As expected, the situation radically changes as the grains ap-
proach the sublimation region. As already identified in previous
studies, as the grains start to sublimate, radiation-pressure in-
creases and eventually halts their inward drift. During this "stage
II" (to follow again Kobayashi et al. 2009), the grain shrinks
while staying at its sublimation radius rs, which does not always
correspond to a constant distance to the star because grain tem-
peratures, and thus their sublimation distance, depend on their
size (see for instance Fig. 2e). In parallel with this size decrease,
the grain's eccentricity increases rapidly. At one point, this ec-
centricity becomes significant enough for the particle to spend
only a very small fraction of its orbit in the narrow sublimat-
ing region around rs. The grain then enters "stage III", where
its sublimation drastically slows down, only occurring at peri-
astron passages. Its orbital eccentricity continues to increase,
albeit more slowly than before, receiving additional "kicks" at
each sublimating-periastron passage.
The final fate of the grain was not investigated in Kobayashi
et al. (2009) or Van Lieshout et al. (2014) because it occurs in
a fast-evolving regime where orbit-averaged equations are no
longer valid. The final fate depends on the dust composition and
the stellar type. This is discussed in details below.
3.1.2. Final fate: ejection
In most cases, the dust grain is parked on these ever-more-
eccentric orbits, all having their periastron at rs, until its eccen-
tricity reaches 1 and the grain is ejected from the system. Note
that, by the time it reaches the e = 1 limit, its β is higher than the
classical 0.5 value expected for a grain released from a β = 0 pro-
genitor on a circular orbit (Fig. 4). This is because, when grains
start to sublimate, in stage II, PR-drag is still able to force their
eccentricities to low values, lower than the one they should have
according to the canonical e = β/(1 − β) relation. So that once
sublimation gets really intensive and the grain approaches the
β = 0.5 value, its eccentricity is still relatively small, allowing it
to stay on a bound orbit beyond this critical 0.5 value. The high-
est possible β value for a grain reaching e = 1 is β = 1, obtained
for an idealized case where the e = 1 grain is produced from
a β = 0.5 progenitor on a circular orbit. However, in practice,
we never obtain β values exceeding 0.8-0.85 (see Fig. 4 for the
simulation leading to the highest beta values for e = 1 particles),
which is because the grains' eccentricities are never exactly zero
as they enter stage II. And even values as small as a few 10−4
are enough to prevent the orbit from reaching β = 1 by the time
it reaches e = 1 (this can be understood by looking at Equ. 58
of Kobayashi et al. (2009) and Equ. 48 of Van Lieshout et al.
(2014), which give the evolution of e during stage II as a func-
tion of the initial e it has when it enters this stage).
Fig. 4: PR-drag pile-up scenario: Evolution of β as a function
of stellar distance r for the initially bound astrosilicate grains
around the A0 star. The parent belt is located at 50 au (vertical
dashed line), and the grains are released on increasingly eccen-
tric orbits as β raises. The arrows represent the inital apoastron
of the grains.
The fact that β < 0.85 by the time the grains are ejected has
important consequences. It means that the DDE always remains
negligible, because its magnitude only becomes significant for β
values very close to 1. Taking Eq. 2 and 3, the ratio of the DDE
force to the radiation pressure+PR-drag force along the velocity
indeed reads :
(cid:112)rGM(cid:63)(1 − β)
ω(cid:63)R2
(cid:63)
FDDE
FPR
=
4
.
(8)
For the highest β value obtained in our runs, i.e., 0.85 for an
F0 star and carbon grains, we get a maximum FDDE/FPR value
of 0.03. We can thus safely conclude that DDE only has a very
marginal influence on the grains' evolution, whose effect can be
neglected on the density pile-up of grains close to rs.
3.1.3. Final fate: total sublimation
For a small subset of our simulations, the grains' fate is radically
different, as they are removed from the system by total subli-
mation. These correspond to the specific cases of a G0 star and
glassy silicates and of a K0 star for both astrosilicate and glassy
silicates (Fig. 3). The behaviour for the K0 cases is easy to un-
derstand: the maximum possible β value is indeed always below
1, regardless of particle sizes (Fig. 1). This means that, as they
sublimate during stage II, grains will never reach β values high
enough for them to reach the e = 1 limit. They will thus stay on
bound orbits all the time, while still sublimating at their orbital
periastron, so that they eventually get fully sublimated. As a con-
sequence, the grains that reach the maximum possible β value
will survive longer than grains close in size but with smaller β
values.
For a G0 star and glassy silicates, β can exceed one, but only
in a relatively narrow size range (see Fig.1c). As a consequence,
as it sublimates, the radius of a grain can directly cross the whole
β > 1 size domain, and even sometimes the β > 0.5 one, before
having had the time to be pushed on an unbound orbit. The fate
Article number, page 7 of 21
10-310-210-1100101102r (au)10-210-1100β1e-05 m2e-05 m3e-05 m5e-05 m1e-04 m2e-04 m3e-04 mA&A proofs: manuscript no. V_6.1
As can be seen in Fig. 6, the τ(r) profiles are almost flat for
most of the domain between the release belt position down to rs,
which is the expected result for a PR-drag scenario (Burns et al.
1979). The only departure from the flat profile occurs close to
the sublimation radius, rs, where we obtain a dust pile-up gener-
ating a density enhancement of a factor of a few at most (earliest
type stars, carbon grains) with respect to the plateau at larger
distance, which is compatible with the values obtained by Van
Lieshout et al. (2014). This enhancement is due to the biggest
grains sublimating at rs, all passing through the (r,s) = (0.2 au,
10−5 m) bin in the 2D maps in the case of an A0 star and car-
bon grains (see also the upper left panel of Fig. 8) before being
expelled. The radial extent of this pile-up is also very narrow,
and is only marginally resolved in our simulations. Thus we can
constrain the enhancement to occur over less than 10−2 au, cor-
responding to a ratio ∆r/r of around 0.1. We note that, for as-
trosilicates and glassy silicates, the pile-up is even weaker than
for carbon grains. This is because, for most stellar types, the β
ratio is below 1 for the smallest grains of these compositions
(Fig. 1). This means that, as grains start to sublimate close to rs,
they will not stay a long time on eccentric orbits (the reason for
the pile-up) before sublimating completely.
3.2.3. SED
As mentioned in the introduction, the best way to evaluate how
well our numerical scenario is able to explain exozodis ob-
servations is not to estimate how it reproduces the predictions
of radiative transfer models regarding dust size or pile-up, but
rather to evaluate how well it reproduces the observational con-
straints themselves. To this effect, we use the Python version of
the GRaTeR code developed for this study to generate synthetic
SEDs. As explained in Sec. 2.1, because we neglect collisional
effects in the parent belt, we cannot constraint the absolute level
of dustiness, and thus the absolute near-IR fluxes. But we can
focus on the relative balance between the near-IR and mid-IR
fluxes as the main criteria to assess the validity of exozodi pro-
ducing scenarios. As a consequence, we chose to rescale all our
synthetic SEDs in order for the emission at 2 µm to correspond
to the level measured for 4 observationally detected exozodis
corresponding to the 4 different spectral types considered: Vega
(A0), η Corvi (F0), 10 Tau (G0) and τ Ceti (a G8 star that is rel-
atively close to a K0 one). We note that, for all these cases, the
flux excess at 2 µm is always of the order of ∼ 1% of the stellar
contribution.
The four corresponding synthetic SEDs are shown in Figs. 7a
to d. We clearly see that, for all considered spectral types and
grain compositions, the shape of the synthetic SED contradicts
the observational constraints. The synthetic SEDs peak in the
far-IR and the flux density in the 10 -- 20 µm domain is always
much higher than the one found in exozodi observations, which
are only of the order of a few percent of the stellar flux around
10 µm. This clearly illustrates the fact that the pile-up near the
sublimation region is far from being sufficient to boost the near-
IR flux at the point where it can dominates over the mid-IR flux.
This mid-IR flux excess is due to the continuous flow of PR-drag
drifting grains in the region between the production belt and rs.
This can be clearly seen in Figure 8, which shows, for the two
"extreme" A0 and K0 cases, the contributions of the grains to
the fluxes at 2 and 10 µm, as a function of their size and spatial
location.
We checked if one way to alleviate this problem could be to
start with a parent belt much closer than the considered 50 au. We
reproduce this situation in a simple manner, without running ad-
Fig. 5: Arrows representing the initial to final grain size evolu-
tion for a sample of simulations. The arrow head corresponds to
the ejection size, or to total sublimation if it reaches the bottom
of the figure. The colored rectangles are the range of grain sizes
for which β > 0.5.
of a grain is very difficult to predict in advance, as it strongly de-
pends on its orbital location by the time it begins to significantly
sublimate. All we can safely establish is that there is a significant
fraction of grains that will disappear because of full sublimation
(Fig. 3a).
3.2. Global disk properties and spectra
3.2.1. Grain size
Figure 5 and Table 3 provide an overview of the smallest grain
sizes that can be reached for our sample of stars and grain com-
positions. We see that, for early-type stars, we are unable to pro-
duce submicron-sized grains regardless of the considered grain
composition. This absence of submicron-sized grains extends
also to the case of late-type stars when considering carbonaceous
dust. These results are in apparent contradiction with constraints
on dominant grain sizes derived from precise spectral modeling
of observed exozodis, which always tend to favour submicron-
sized dust (see Sec. 1).
The only cases for which submicron-sized grains are pro-
duced are those where the grain can experience full sublimation,
i.e., those for which the maximum possible β value is lower than
1 or barely exceeds it. As has been discussed in Sec. 3.1.3, this
is only true for K0 stars (astro- and glassy silicates) and G0 stars
(glassy silicates only). However, even in these cases, the lifetime
of such tiny grains is very short (sublimation being very fast and
efficient), which might not be enough to leave an observable sig-
nature.
3.2.2. Surface density profiles
Another prediction of radiative transfer models of exozodis is
that the hot dust is expected to be confined close to its sublima-
tion radius. In order to evaluate the level of dust pile-up at rs, we
compute the radial distribution of the total geometrical optical
depth, τ, of the dust produced by the PR-drag scenario, follow-
ing the approach described in App. C.
Article number, page 8 of 21
A0F0G0K0Spectral type10-910-810-710-610-510-410-3s (m)GlassySiAstroSiCarbonÉ. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(a) Carbon
(b) Astrosilicate
(c) Glassy silicate
Fig. 6: PR-drag pile-up scenario: Radial profiles for the (size-integrated) geometrical optical depth for all grain compositions and
stellar types.
Table 3: PR-drag pile-up scenario: Minimal distance and grain size reached for each configuration. A grain size of 0 means that the
grain is fully sublimated. If two values are indicated, they correspond to the final sizes for the smallest and biggest grains.
A0 star
F0 star
G0 star
K0 star
rmin (au)
smin (µm)
rmin (au)
smin (µm)
rmin (au)
smin (µm)
rmin (au)
smin (µm)
Carbon
Astrosilicate
Glassy silicate
0.16
0.44
0.38
7.6
4.6
3.1
0.05
0.14
0.07
1.5
0 / 0.9
0
0.03
0.08
0.02
0.9
0 / 0.5
0
0.02
0.04
0.01
0.4
0
0
ditional simulations. We consider our original simulations with
a parent belt at 50 au, and integrate the flux coming from the
grains within a given distance to the star. That distance is as-
sumed to mimic the new location of the parent belt. In the spe-
cific case of Vega, the results are shown Fig. 7e, where all fluxes
have been normalized to the same observed flux at λ = 2.12 µm.
The "unwanted" mid-IR (10 µm) flux falls down to observation-
compatible levels only for an extremely close-in parent belt lo-
cated at 0.4 au. This solution appears highly unlikely given that
such a massive collisional belt would probably not be able to
survive long enough so close to the star to sustain the hot dust
for a duration comparable to the age of the star.
Overall, these conclusions hint at another production process
than the PR-drag mechanism to populate the hot exozodiacal
dust systems.
4. Exocometary dust delivery scenario
Another classic process of exozodiacal dust production is the
cometary grain release very close to the star. In this scenario, an
outer mass reservoir remains necessary, but the dust grains are
deposited by large, undetectable parent bodies in the immediate
vicinity of the place where they are detected. A benefit of this
process compared to the PR-drag pile-up scenario is that it leaves
essentially no observable signature in between the parent belt
and the exozodi.
In Bonsor et al. (2012), we investigated the planetary sys-
tem architecture required to sustain an inwards flux of exo-
comets. In Bonsor et al. (2013), we highlighted the importance
of planetesimal-driven migration of the planet closest to the in-
ner edge of the belt to maintain this flux on sufficiently long
timescales (see also Raymond & Bonsor 2014). In Marboeuf
et al. (2016), we evaluated the cometary dust ejection rate as a
function of the distance to the star and spectral type, to help con-
necting dynamical simulations to exozodi observations in future
studies (e.g. Faramaz et al. 2017). Here, we make an important
step further by discussing the fate of the grains once released
by an exocomet passing close to the star, and by calculating the
resulting emission spectra for a direct comparison to the data.
4.1. Numerical setup
In order to compare the outcome of our cometary model to the
results obtained for the PR-drag pile-up scenario (Sec. 3), we
consider a reservoir of exocomets that have their aphelion at a
fixed distance of 50 au and a perihelion rp just outside the subli-
mation limit, which, for each composition and spectral type, we
define as the largest sublimation distance of the considered grain
sizes (often corresponding to the smallest grains, see e.g. vertical
dashed lines in Fig. 9a to 9d).
We assume that all grains leaving the comet are produced
when the comet passes at perihelion rp. This is a simplifying
assumption because grains should be dragged from the comet
by the evaporation of volatiles, which should happen over a
large fraction of its orbit. However, as shown by Marboeuf et al.
(2016), the volatile and dust production rate strongly increase
with decreasing distance to the star (see Equ.17 of that paper),
so that most of the mass loss happens in a narrow region close to
the comet's perihelion, as is clearly illustrated by Fig.10.
Grains produced at perihelion have the highest possible
speed once released from the comet and are thus less prone to
remain bound. More precisely, the blowout limit in term of β is
lowered compared to the PR-drag pile-up scenario and can be
expressed using (e.g. Murray & Dermott 1999):
βblow =
1
2
1 − e2
0
1 + e0 cos Φ
.
(9)
where Φ is the longitude of the release position on the cometary
orbit, and e0 the parent body eccentricity. This reduces to βblow =
(1 − e0)/2 for a release at perihelion. For the grain compositions
and spectral types explored in this study (Sec. 2.3), the release
distance varies between 2 au and ∼ 0.02 au, corresponding, for
Article number, page 9 of 21
10-210-1100r (au)0.00.51.01.52.0τgeo normalizedA0F0G0K010-210-1100r (au)0.00.51.01.52.0τgeo normalizedA0F0G0K010-210-1100r (au)0.00.51.01.52.0τgeo normalizedA0F0G0K0A&A proofs: manuscript no. V_6.1
(a)
(c)
(b)
(d)
(e)
Fig. 7: Panels (a) to (d) : SED for each spectral type and all compositions in the PR-drag pile-up scenario. All spectra are normalized
to the flux ratio at 2 µm, and compared to observed fluxes for exozodis around stars of similar type (for the F0 star, for which there
is no available observed photometry at 2 µm, the value is put to 1% of the stellar flux). Data points for Vega are from Absil et al.
(2006), η Crv are from Lebreton et al. (2016), 10 Tau are from Kirchschlager et al. (2017), τ Cet are from Di Folco et al. (2007).
Additional values are from Absil et al. (2013) and Mennesson et al. (2014). Panel (e): SEDs for an A0 star and carbon grains, but
only considering the flux within a radial distance indicated in the top-right corner, and flux-normalized at λ = 2.12 µm such that the
disk flux amounts to 1.29% of the stellar flux at that wavelength (FLUOR excess for Vega).
Article number, page 10 of 21
10-1100101102103Wavelength (µm)10-1100101102103104Flux density (Jy)A0 starCarbonAstrosilicateGlassy silicateVega10-1100101102103Wavelength (µm)10-310-210-1100101102Flux density (Jy)F0 starCarbonAstrosilicateGlassy silicateη Crv10-1100101102103Wavelength (µm)10-310-210-1100101102Flux density (Jy)G0 starCarbonAstrosilicateGlassy silicate10 Tau10-1100101102103Wavelength (µm)10-310-210-1100101102Flux density (Jy)K0 starCarbonAstrosilicateGlassy silicateτ Cet10-1100101102103λ (µm)10-1100101102103104SED (Jy)Star0.4 au0.9 au2.2 au5.4 au13.0 au31.5 auÉ. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(a)
(c)
(b)
(d)
Fig. 8: PR-drag pile-up scenario: Flux level at 2 µm (left) and 10 µm (right), for an A0 star and carbon grains (top); and K0 star and
astrosilicate grains (bottom). See Sec. 2.4 and App. C for the methodology. The verticalish dashed line correspond to the location of
the sublimation radius rs as a function of grain size, while the full and dashed horizontal lines indicate the β = 0.5 and β = 1 limits,
respectively.
an apoastron of 50 au, to parent body eccentricities varying be-
tween 0.923 and 0.999, and βblow values between 3.8 × 10−2 and
5.0 × 10−4, respectively. These low βblow values translate into
large grain sizes, of several tens to several hundred of µm, for the
limiting blow-out size. The release distances, orbit eccentricities,
blowout β values and grain sizes sblow are documented in Tab. 4
for the four spectral types and three compositions investigated in
this study.
4.2. Grain evolution
4.2.1. Carbon and astrosilicate grains
In our model, the carbon and astrosilicate bound grains (β <
βblow, i.e. s > sblow) are delivered by an exocomet a little beyond
their sublimation distance, leaving room for dynamical evolu-
tion. The fate of these grains shows similarities with that dis-
cussed in the case of the PR-drag pile-up scenario. Their semi-
major axis and eccentricity both decrease by PR-drag, until they
are totally sublimated (Sec. 3.1.3) or expelled from the system
by radiation pressure due to partial sublimation (Sec. 3.1.2), as
illustrated in Figs. 9c and d.
There are, however,
important differences between the
cometary and PR-drag pile-up scenarios. In particular, the high
orbital eccentricity of the grains inherited from the comet im-
plies that those grains spend only a very small fraction of their
orbital period (typically less than a day) close to the sublima-
tion zone in the early phases of their evolution. Therefore, these
grains see their size remaining essentially constant while migrat-
ing inward by PR-drag, like in the stage I of the PR-drag pile-up
scenario (Sec. 3.1.1). Their orbit is getting circularized until sub-
limation becomes significant, thereby moving into the stage II
phase. However, they enter this phase with a significant residual
eccentricity (of the order of 0.1, e.g. Fig. 9e), much larger than in
the case of the PR-drag pile-up scenario. As a consequence, the
carbon and astrosilicate grains are expelled when they reach a β
value of about 0.6 (Fig. 9f), to be compared to the 0.8 accessible
in the PR-drag scenario. The only exception is for astrosilicate
grains around the K0-type star, for which the β = 0.6 value is
never reached accross all grain sizes, meaning that the bound
Article number, page 11 of 21
1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)8.07.26.45.64.84.03.22.41.6Log flux at 2µm (Jy)1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)87654321Log flux at 10µm (Jy)1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)1816141210864Log flux at 2µm (Jy)1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)161412108642Log flux at 10µm (Jy)A&A proofs: manuscript no. V_6.1
Table 4: Parameters for the cometary release at perihelion scenario : assumed release position (rp, in au) and resulting comet
orbit eccentricity (e), blowout β value (βblow) and corresponding blowout grain size (sblow, in µm) (see Sec. 4.1 for details on the
assumptions).
Carbon
Astrosilicate
Glassy silicate
Carbon
Astrosilicate
Glassy silicate
rp
0.60
2.4
1.6
rp
0.081
0.23
0.050
0.976
0.908
0.938
e
e
0.997
0.991
0.998
A0 star
G0 star
βblow
1.2×10−2
4.6×10−2
3.1×10−2
βblow
1.6×10−3
4.5×10−3
9.9×10−4
sblow
470
70
90
sblow
330
65
230
rp
0.16
0.43
0.11
rp
0.038
0.11
0.027
e
0.994
0.983
0.996
F0 star
βblow
3.1×10−3
8.5×10−3
2.1×10−2
K0 star
e
0.999
0.996
0.999
βblow
7.6×10−4
2.1×10−3
5.3×10−4
sblow
310
65
210
sblow
240
51
120
(a) A0, glassy silicate
(b) G0, glassy silicate
(c) K0, astrosilicate
(d) F0 star, Carbon
(e) F0 star, Carbon
(f) F0 star, Carbon
Fig. 9: Panels (a) to (c) : grain size as function of the distance for all simulations ran in the exocometary dust delivery at perihelion
scenario (the yellow area, the dotted and dash-dotted lines have the same meaning as in Fig. 2, while dashed line corresponds to the
limit between initially bound and unbound grains sizes). Panels (d) to (f) : specific case of an F0 star and carbon grains, for which
the evolution of the grain size, the eccentricity and the β value as a function of distance is displayed. In the case of eccentricity, the
two curves overlap.
grains end up completely sublimated. Another consequence of
the significant residual eccentricity during stage II is that, here
again, the grains do not reach β values close enough to 1 for the
DDE mechanism to operate.
4.2.2. Glassy silicate grains
The behaviour of glassy silicate grains is quite different. The
sublimation timescale of these grains at the sublimation temper-
ature is four orders of magnitude lower than for the other com-
positions considered here (Sec.2.3). In this case, the sublimation
time becomes comparable to the time spent by the grain close to
the sublimation limit in only one perihelion flyby. For the A- and
F-type stars, this results in a complete sublimation of the bound
grains after the circularization of stage II.
Article number, page 12 of 21
For later type stars, the fate of the bound grains is affected
by the unusual fact that, at a given distance, large glassy silicate
grains are hotter than smaller ones (see e.g. the almost vertical
dash-dotted lines in Figs. 3 and 9b). Therefore, the large bound
glassy silicate grains are released at, or very close to, their subli-
mation distance around late-type stars in our model. Their high
temperature, combined with the intrinsic sublimation efficiency
of glassy silicates compared to carbon and astrosilicate grains,
implies that their sublimation timescale is in this case lower than
the PR-drag timescale. As consequence, these bound grains be-
come small enough to be expelled before their orbits can be cir-
cularized.
10-310-210-1100101102r (au)10-810-710-610-510-410-3s (m)A0 star, Glassy Silicate10-310-210-1100101102r (au)10-810-710-610-510-410-3s (m)G0 star, Glassy Silicate10-310-210-1100101102r (au)10-910-810-710-610-510-410-3s (m)K0 star, Astrosilicate10-310-210-1100101102r (au)10-910-810-710-610-510-410-3s (m)F0 star, Carbon10-310-210-1100101102Semi-major axis (au)10-210-1100eF0 star, Carbon5e-04 m1e-03 m10-310-210-1100101102r (au)10-410-310-210-1100β2e-06 m3e-06 m5e-06 m1e-05 m2e-05 m3e-05 m5e-05 m1e-04 m2e-04 m3e-04 m5e-04 m1e-03 mÉ. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
As can be seen in Fig. 11, we obtain a flat τ(r) profile in between
rs and 1 au, followed by a r−1.7 profile at larger distances. The
plateau between rp to rs results from the inward migration and
circularization of the orbit of the bound grains by PR-drag, as
already observed for the PR-drag pile-up scenario (Fig. 6), while
the high-eccentricity bound grains populate the regions outside
that distance. We conclude that the exocometary dust delivery
position can essentially be regarded as playing the same role as
the parent belt distance in the PR-drag pile-up scenario, as far
as the shape of the τ(r) profile between that reference position
down to the sublimation distance is concerned. Numerically, the
PR-drag pile-up scenario can be considered as an extreme case
of the cometary scenario, with exocomets having zero eccentric-
ity.
Fig. 10: Fraction of the surfacic mass loss for an exocomet or-
biting an F0 star as a function of radial distance. The comet has
an apoastron of 50 au, and an eccentricity of 0.994. Based on the
comet evaporation prescription of Marboeuf et al. (2016, their
Eq. 17).
Fig. 11: Total geometrical optical depth (τ) in the exocometary
dust delivery at perihelion scenario, assuming carbon grains. For
illustrative purpose, a model with an exocomet perihelion larger
than in the nominal case is shown for the F0 star.
4.3. Global disk properties and spectra
4.3.1. Surface density profiles
We calculate the radial distribution of the total geometrical opti-
cal depth, τ, of the dust produced in the comet-release scenario,
following the approach described in App. C, and already used in
the case of the PR-drag pile-up scenario (see Sec. 3.2.2). As il-
lustrated in Fig. 11 for carbon grains, for a release at a perihelion
rp that is close to the sublimation distance rs, the radial density
profile peaks at the release location and decreases further out
with the distance to the star as r−1.7.
To check the importance of the position of the periastron
with respect to the sublimation distance, we performed an ad-
ditional simulation, for an F0 star and carbon grains, for which
the dust is released at 1 au instead of 0.13 au in the nominal case.
4.3.2. SED
We use the same computing and normalization approach as in
Sec. 3.2.3 to evaluate the SEDs resulting from the exocometary
dust production scenario. A notable difference with the PR-drag
pile-up scenario, is that it is possible here to connect the absolute
exozodiacal dust level to a number of exocomets passing close
to the star, as discussed further in Sec. 4.4.
The synthetic SEDs for the exocometary scenario are dis-
played in Fig. 12 for the four spectral types and three dust com-
positions considered in this study, and are compared to the same
reference observed systems as in Fig. 7. Overall, the mid- to far-
infrared excess is significantly reduced compared to the PR-drag
pile-up scenario. This is mainly because the grains are directly
deposited next to the sublimation zone, mitigating the amount of
dust beyond the grain release position (comet's perihelion). The
peak of the SED is accordingly shifted to much shorter wave-
lengths compared to the PR-drag pile-up scenario.
In this respect, the carbon grain case is the most favourable
one. Regardless of the spectral type, the exozodiacal emission
of cometary carbon grains peaks at 3 -- 5 µm. The flux at 10 µm
is always significantly smaller than the stellar flux at the same
wavelength, by a factor of about five, even though it is still not
enough to fully fit the data. The models with astrosilicate and
glassy silicate grains, on the other hand, predict much too large
emissions in the mid- and far-infrared compared to the observa-
tions of hot exozodis.
In Fig. 12b we also display the SED obtained for the case
with a comet perihelion at rp = 1 au, far away from the subli-
mation distance rs. As can be clearly seen, the fit to the observed
data is much poorer because of the excess mid-IR flux due to PR-
drag drifting grains in the flat region between rp and rs that was
identified in Fig.11. In essence, we are here close to the cases
explored in the PR-drag pile-up scenario (Sec.3).
The main contributors to the flux at different wavelengths can
be explored using 2D emission maps as a function of the grain
size and distance to the star. Figures 13a to d show examples for
our best case, i.e., an F0 star and a carbon dust composition. We
see that the near- and mid-IR emissions essentially come from
the same relatively large grains (several tens to hundreds of mi-
crometers), which are the smallest bound grains released by the
exocomets (Tab. 4). At 2 µm, most of the flux originates from
the smallest bound grains just outside the sublimation distance,
while the 10 µm emission comes from similar grains in size but
distributed in a broader region centered around the dust release
position (Fig. 13b). At that wavelength, the drop of the surface
density beyond the dust delivery position (0.13 au) contributes to
moderating the emission from the distant regions, as was demon-
strated in the PR-drag pile-up scenario when we schematically
Article number, page 13 of 21
10-1100101102r (au)0.000.050.100.150.200.250.300.35Distribution of mass production10-210-1100101r (au)10-1010-910-810-710-6τ integratedA0F0, 0.13 auF0, 1.0 auG0K0A&A proofs: manuscript no. V_6.1
(a)
(c)
(b)
(d)
Fig. 12: Same as Fig. 7 but for the exocometary dust delivery at perihelion scenario.
simulated much closer-in parent belts (Fig. 7). These behaviours
at 2 and 10 µm are emphasized when the exocomet's perihelion
is arbitrarily moved to 1 au, as shown in Fig.13c and d. Finally,
we note that a direct consequence of the fact that the emission
originates from large grains is that the strong silicate features
seen in case of the PR-drag pile-up scenario are here essentially
absent.
Overall, it is noteworthy that two key features of the classical
radiative transfer models of exozodis (e.g. Absil et al. 2006, and
following studies) are reproduced with the exocometary dust de-
livery scenario, namely an accumulation of the grains very close
to the sublimation zone and a preference for carbon-rich dust.
Nevertheless, it should also be noted that the carriers of the exo-
zodi emission in that case are several orders of magnitude larger
in size than the grains usually required by the classical radiative
transfer models.
Article number, page 14 of 21
4.4. Inward flux and size of the exocomets
As already mentioned, in the previous sections the flux was ar-
bitrarily rescaled in order to match the observed excess levels in
the near-IR. Now, in order to discuss further the relevance of the
exocometary dust delivery scenario, we aim to evaluate the size
and number of exocomets required to physically reach these ob-
served exozodiacal flux level. For that purpose, we employ the
cometary dust ejection prescription of Marboeuf et al. (2016), in
particular their equation 17, together with the orbital parameters
documented in Table 4, to quantify the mass of grains released
by a comet over an orbital period. This also allows to estimate,
for a given comet mass, the exocomet lifetime as well as the
number of orbits before complete erosion. These quantities are
used to assess the absolute flux density at 2 and 10 µm result-
ing from the evaporation of an exocomet of a certain size. The
different steps of the adopted methodology are detailled in Ap-
pendix D, and the results for a typical 10 km-sized exocomet are
summarized in Tables D.1 and D.2.
In the case of carbon and astrosilicate grains, the flux pro-
duced at 2 µm by one 10 km-sized exocomet amounts to a few
10-1100101102103Wavelength (µm)10-1100101102103104Flux density (Jy)A0 starCarbonAstrosilicateGlassy silicateVega10-1100101102103Wavelength (µm)10-310-210-1100101102Flux density (Jy)F0 starCarbonAstrosilicateGlassy silicateCarbon, 1 auη Crv10-1100101102103Wavelength (µm)10-310-210-1100101102Flux density (Jy)G0 starCarbonAstrosilicateGlassy silicate10 Tau10-1100101102103Wavelength (µm)10-310-210-1100101Flux density (Jy)K0 starCarbonAstrosilicateGlassy silicateτ CetÉ. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
(a)
(c)
(b)
(d)
Fig. 13: Contribution of each size and distance bin to the flux at 2 µm (left) and 10 µm (right) in the case of an F0 star and carbon
grains. The grains are released by a exocomet at 0.13 au (top) and 1 au (bottom).
10−5 to a few 10−4 the stellar flux at the same wavelength. There-
fore, a few tens to a few hundred of active 10 km-radius exo-
comets on a similar orbit are required to reach the observed level
of ∼1% of the stellar flux at 2 µm. At 10 µm, the dust to star flux
ratio is always larger than at 2 µm, by a factor of about twenty for
the carbon grains, and a factor of about 200 for the astrosilicate
grains, in agreement with the results in Sec. 4.3.2. The required
number of active exocomets can be mitigated if their initial size
is larger. Indeed, because we consider the total flux resulting
from the complete erosion of the exocomet, it scales directly
with the exocomet mass and hence with the exocomet radius to
the cube. Therefore, a single active 40 to 80 km-sized exocomet
would be enough to produce a 1% excess at 2 µm. Glassy sili-
cates fluxes are significantly lower due to the short sublimation
lifetime of such grains. Around FGK stars, these grains subli-
mate significantly at each perihelion passage, and disapear in a
few orbits.
Several assumptions enter into the calculation of the num-
ber and size of the exocomets required to reproduce the obser-
vations (see Sec. 4.1 and App. D), and the above values should
therefore be taken with caution. It is also useful to remind that
our conclusions are valid for the specific orbital parameters sum-
marized in Table 4. Nevertheless, the results are encouraging in
the sense that the sizes and number of exocomets appear reason-
able, thereby providing additional support to the exocometary
dust delivery scenario explored in this study.
5. Summary and conclusion
By investigating, with non orbit-averaged equations, the fate
of dust around several different stellar types, and considering
different grain compositions, we are able to draw some con-
clusions on the properties of exozodis emission depending if
these grains come from an outer parent belt and drift inward by
PR-drag (Sec.3), or if they have an exocometary origin (Sec. 4).
We show that :
• in the case of the PR-drag pile-up scenario :
-- for early-type stars, significant amounts of sub- µm sized
grains cannot be produced in the inner disc regions, because
grains are blown-out by radiation pressure before sublimat-
ing down to these sizes. The only case for which sub- µm
Article number, page 15 of 21
1e-021e-011e+001e+01r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)18.016.515.013.512.010.59.07.56.0Log flux at 2µm (Jy)1e-021e-011e+001e+01r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)181614121086Log flux at 10µm (Jy)1e-021e-011e+001e+01r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)16.515.013.512.010.59.07.56.0Log flux at 2µm (Jy)1e-021e-011e+001e+01r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)18.016.515.013.512.010.59.07.56.0Log flux at 10µm (Jy)A&A proofs: manuscript no. V_6.1
grains are obtained is that of silicate grains around later-type
stars,
-- dust pile-up close to the sublimation radius is moderate, gen-
erating a density enhancement of a few at most with respect
to an otherwise flat surface density profile,
-- the near-IR excess is always associated to mid-IR excess at
the same level, or even much higher. This latest behaviour
cannot explain the numerous near-IR excesses without mid-
IR excess detections.
• in the case of the Exocometary dust delivery scenario :
-- a narrow ring forms close to the sublimation zone, near
the comet's periastron. This ring is predominantly populated
with large grains, a few tens to a few hundred µm in radius
depending on the composition, which are the smallest bound
grains produced at the comet's perihelion,
-- compared to the PR-drag pile-up scenario, the near-IR excess
is associated with a much smaller mid-IR excess, in better
agreement with the data, although not totally fitting them.
Carbon-rich grains provide the best results,
-- the near-IR excess can be reproduced assuming a realistic
inward fluxes of exocomets and reasonable exocomet sizes.
In addition, we find that the DDE mechanism, which could in
principle help forming a dense dust ring close to the sublimation
radius, has a very limited effect in both scenarios, because parti-
cles never reach β values close enough to 1 by the time they are
blown out.
Therefore, based on simulations performed with the new
numerical model developed in the context of this study, we
conclude that the PR-drag pile-up scenario is unlikely to pro-
duce the hot exozodis observed with near-IR interferometry. The
exocomet-release at perihelion scenario, on the other hand, pro-
vides a very promising theoretical framework that should be ex-
plored further. For that purpose, future developments of our nu-
merical model shall include a post-processing treatment of the
collisions in the dust ring close to the sublimation zone, build-
ing up on the DyCoSS collisional model used in the context of
debris disks (e.g. Thébault et al. 2012, 2014). It should also in-
clude a detailed calculation of the grain charging, following for
instance the prescription of Kimura et al. (2018), and a careful
consideration of the magnetic topology, in order to thoroughly
discuss the efficiency of magnetic trapping and its impact on the
near- and mid-IR emissions.
Acknowledgements. We thank Hiroshi Kimura for providing the optical con-
stants for glassy silicates, and for the discussions about the DDE mechanism. We
also thank Rik van Lieshout for useful discussion concerning sublimation pro-
cesses. We acknowledge the financial support from the Programme National de
Planétologie (PNP) of CNRS-INSU co-funded by the CNES. Our code uses the
NumPy (Oliphant 2006) and the Matplotlib libraries (Hunter 2007). This work
has made use of data from the European Space Agency (ESA) mission Gaia
(https://www.cosmos.esa.int/gaia), processed by the Gaia Data Process-
ing and Analysis Consortium (DPAC, https://www.cosmos.esa.int/web/
gaia/dpac/consortium). Funding for the DPAC has been provided by na-
tional institutions, in particular the institutions participating in the Gaia Mul-
tilateral Agreement.
References
Absil, O., di Folco, E., Mérand, A., et al. 2006, Astronomy &
Akeson, R. L., Ciardi, D. R., Millan-Gabet, R., et al. 2009, The
Astrophysical Journal, 691, 1896
Augereau, J.-C., Lagrange, A.-M., Mouillet, D., Papaloizou, J.
C. B., & Grorod, P. A. 1999, Astronomy & Astrophysics, 348,
557
Belton, M. J. S. 1966, Science, 151, 35
Beust, H. & Morbidelli, A. 2000, Icarus, 143, 170
Bonsor, A., Augereau, J.-C., & Thébault, P. 2012, Astronomy &
Astrophysics, 548, A104
Bonsor, A., Raymond, S. N., & Augereau, J.-C. 2013, Monthly
Notices of the Royal Astronomical Society, 433, 2938
Bonsor, A., Raymond, S. N., Augereau, J.-C., & Ormel, C. W.
2014, Monthly Notices of the Royal Astronomical Society,
441, 2380
Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1
Cameron, A. G. W. & Fegley, M. B. 1982, Icarus, 52, 1
Czechowski, A. & Mann, I. 2012, in Astrophysics and Space
Science Library, Vol. 385, Nanodust in the Solar System: Dis-
coveries and Interpretations, ed. I. Mann, N. Meyer-Vernet, &
A. Czechowski, 47
Defrère, D., Absil, O., Augereau, J.-C., et al. 2011, Astronomy
& Astrophysics, 534, A5
Dermott, S. F., Jayaraman, S., Xu, Y. L., Grogan, K., &
Gustafson, B. A. S. 1996, in American Institute of Physics
Conference Series, Vol. 348, American Institute of Physics
Conference Series, ed. E. Dwek, 25 -- 36
Di Folco, E., Absil, O., Augereau, J.-C., et al. 2007, Astronomy
& Astrophysics, 475, 243
Draine, B. T. 2003, The Astrophysical Journal, 598, 1026
Ducati, J. R. 2002, VizieR Online Data Catalog
Ertel, S., Absil, O., Defrère, D., et al. 2014, Astronomy & As-
Ertel, S., Defrère, D., Absil, O., et al. 2016, Astronomy & As-
trophysics, 570, A128
trophysics, 595, A44
Ertel, S., Defrère, D., Hinz, P. M., et al. 2018, The Astronomical
Journal, 155 [arXiv:1803.11265]
Faramaz, V., Ertel, S., Booth, M., Cuadra, J., & Simmonds, C.
2017, Monthly Notices of the Royal Astronomical Society,
465, 2352
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018,
Astronomy & Astrophysics [arXiv:arXiv:1804.09365v2]
Gaia Collaboration, Prusti, T., de Bruijne, J. H. J., et al. 2016,
Astronomy & Astrophysics, 595, 1
Hunter, J. D. 2007, Computing in Science & Engineering, 9, 90
Kama, M., Min, M., & Dominik, C. 2009, Astronomy & Astro-
physics, 506, 1199
Kiefer, F., Lecavelier des Etangs, A., Augereau, J.-C., et al.
2014a, Astronomy & Astrophysics, 561, L10
Kiefer, F., Lecavelier des Etangs, A., Boissier, J., et al. 2014b,
Nature, 514, 462
Kiefer, F., Lecavelier des Etangs, A., Vidal-Madjar, A., et al.
2017, Astronomy & Astrophysics, 132 [arXiv:1709.01732]
Kimura, H., Ishimoto, H., & Mukai, T. 1997, Astronomy & As-
trophysics, 326, 263
and Space Science
Kimura, H., Kunitomo, M., Suzuki, T. K., et al. 2018, Planetary
Kirchschlager, F., Wolf, S., Krivov, A. V., Mutschke, H., & Brun-
ngräber, R. 2017, Monthly Notices of the Royal Astronomical
Society, 467, 1614
Kobayashi, H., Kimura, H., Watanabe, S.-i., Yamamoto, T., &
Müller, S. 2011, Earth, Planet, and Space, 63, 1067
Absil, O., di Folco, E., Mérand, A., et al. 2008, Astronomy &
Kobayashi, H., Watanabe, S.-i., Kimura, H., & Yamamoto, T.
Absil, O., Kervella, P., Mollier, B., et al. 2013, Astronomy &
Kobayashi, H., Watanabe, S.-i., Kimura, H., & Yamamoto, T.
2008, Icarus, 195, 871
2009, Icarus, 201, 395
Astrophysics, 452, 237
Astrophysics, 487, 14
Astrophysics, 104, 1
Article number, page 16 of 21
É. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
Kral, Q., Matrà, L., Wyatt, M. C., & Kennedy, G. M. 2017,
Monthly Notices of the Royal Astronomical Society, 550, 521
Krivov, A. V. 2010, Research in Astronomy and Astrophysics,
10, 383
Krivov, A. V., Kimura, H., & Mann, I. 1998, Icarus, 134, 311
Lamy, P. L. 1974, Astronomy & Astrophysics, 35, 197
Lamy, P. L. 1978, Icarus, 34, 68
Lebreton, J., Beichman, C. A., Bryden, G., et al. 2016, The As-
trophysical Journal, 817, 165
Lebreton, J., van Lieshout, R., Augereau, J.-C., et al. 2013, As-
tronomy & Astrophysics, 555, A146
Liou, J. C., Dermott, S. F., & Xu, Y. L. 1995, Planetary and Space
Science, 43, 717
Marboeuf, U., Bonsor, A., & Augereau, J.-C. 2016, Planetary
and Space Science, 133, 47
Mennesson, B., Absil, O., Lebreton, J., et al. 2013, The Astro-
physical Journal, 763 [arXiv:1211.7143]
Mennesson, B., Millan-Gabet, R., Serabyn, E., et al. 2014, The
Montesinos, B., Eiroa, C., Krivov, A. V., et al. 2016, Astronomy
Astrophysical Journal, 797
& Astrophysics, 51, 31
Mukai, T. & Yamamoto, T. 1979, Publication of the Astronomi-
cal Society of Japan, 31, 585
Mukai, T., Yamamoto, T., Hasegawa, H., Fujiwara, A., & Koike,
C. 1974, Publication of the Astronomical Society of Japan,
26, 445
Murray, C. D. & Dermott, S. F. 1999, Solar system dynamics
Nesvorný, D., Jenniskens, P., Levison, H. F., et al. 2010, The
Astrophysical Journal, 713, 816
Nuñez, P. D., Scott, N. J., Mennesson, B., et al. 2017, Astronomy
& Astrophysics, 113, 1
Oliphant, T. E. 2006, USA: Trelgol Publishing
Pollack, J. B., Toon, O. B., & Khare, B. N. 1973, Icarus, 19, 372
Rappaport, S., Vanderburg, A., Jacobs, T., et al. 2018, Monthly
Notices of the Royal Astronomical Society, 474, 1453
Raymond, S. N. & Bonsor, A. 2014, Monthly Notices of the
Royal Astronomical Society, 442, 18
Rieke, G. H., Gáspár, A., & Ballering, N. P. 2016, The Astro-
physical Journal, 816, 50
Sezestre, É., Augereau, J.-C., Boccaletti, A., & Thébault, P.
2017, Astronomy & Astrophysics, 65, 1
Sibthorpe, B., Kennedy, G. M., Wyatt, M. C., et al. 2018,
Monthly Notices of the Royal Astronomical Society, 475,
3046
Su, K. Y. L., Rieke, G. H., Malhotra, R., et al. 2013, The Astro-
physical Journal, 763 [arXiv:1301.1331]
Thébault, P., Kral, Q., & Augereau, J.-C. 2014, Astronomy &
Thébault, P., Kral, Q., & Ertel, S. 2012, Astronomy & Astro-
Astrophysics, 561, A16
physics, 547, A92
Thureau, N. D., Greaves, J. S., Matthews, B. C., et al. 2014,
Monthly Notices of the Royal Astronomical Society, 445,
2558
van Belle, G. T. & von Braun, K. 2009, The Astrophysical Jour-
nal, 694, 1085
van Leeuwen, F. 2007, Astronomy & Astrophysics, 474, 653
Van Lieshout, R., Dominik, C., Kama, M., & Min, M. 2014, As-
tronomy & Astrophysics, 571, A51
Wyatt, S. P. & Whipple, F. L. 1950, The Astrophysical Journal,
111, 134
Physics, 59, 2966
Zavitsanos, P. D. & Carlson, G. A. 1973, Journal of Chemical
Zubko, V. G., Mennella, V., Colangeli, L., & Bussoletti, E. 1996,
Monthly Notices of the Royal Astronomical Society, 282,
1321
Article number, page 17 of 21
Appendix A: Reproducing Krivov et al. (1998)
A&A proofs: manuscript no. V_6.1
Fig. C.1: Cumulative time map derived from the simulations for
the A0, carbon case.
-- (AC, BC) parameters in Cameron & Fegley (1982):
A =
B =
104
BC
AC
BC
+ 6 − log10
kB
µmu
-- (A, BL) parameters in Lamy (1974):
B = BL − log10
kB
µmu × 1.33322 · 103
(B.5)
(B.6)
(B.7)
Appendix C: Geometrical optical depth map
computation
Our goal is to produce density and optical depth maps from tra-
jectories independently computed for individual grains with our
dynamical code (that includes sublimation). In a first step, we de-
fine a 2D grid of logarithmically-spaced distances to the star (r)
and grain sizes (s). Then, for each single-grain size simulation,
we sum up the times spent by the grain during its lifetime in each
bin of the 2D grid, correcting this time by the initial differential
size distribution assumed to be proportional to s−3.5 (collisional
erosion). This yields a first 2D map of cumulative times, that
is proportional to a density map if the system is assumed to be
at steady-state. As illustrated in figure C.1, the limited number
of grain sizes for which the dynamics has been computed leaves,
however, empty grain size bins in the 2D map (empty lines). This
leads us to develop a complementary approach to fill the holes
in this map.
We define a second 2D (r,s) map, of the same size and same
bin values as the first 2D map. We then estimate three timescales
for each bin in the 2D (r,s) map, to qualitatively evaluate the abil-
ity of a grain to move to a nearby bin due to sublimation, ejection
(radiation pressure) or inward migration (PR-drag). The subli-
mation timescale to move from (ri,si) to (ri,si−1) and the ejection
timescale to move from (ri,si) to (ri+1,si) are both taken from Le-
breton et al. (2013). The PR-drag timescale to move from (ri,si)
to (ri−1,si) is taken from Burns et al. (1979). In each bin, the
shortest timescale gives the dominant physical process and this
is used to predict the trajectory of a grain in the 2D map, allowing
a jump from position (ri,si) to (ri−1,si−1) when sublimation and
Fig. A.1: Top : evolution of the grain size as a function of the
distance to the Sun for carbon grains of initial sizes: 1; 3; 10;
and 30 µm. Bottom : same for the glassy silicate. These plots
reproduce the results presented in Fig. 5 of Krivov et al. (1998).
Appendix B: Thermodynamical properties
The literature is rich in 2-parameter formula for describing the
sublimation process of dust grains, which are in essence simi-
lar but with different notations, bringing some confusion. Here,
we propose a summary of the equations to transform the two
parameters in four different papers in the literature to the (A, B)
parameters of Lebreton et al. (2013) used in this study:
-- (AZ, MZ) parameters in Zavitsanos & Carlson (1973), table 3:
(B.1)
A = MZ
B = AZ + 6 + log10
The value of 6 in the expression of B, coming from the unit
system, is lacking in Lebreton et al. (2013), explaining the
difference with our derived values.
kB
µmu
(B.2)
-- (H, P) parameters in Kobayashi et al. (2009):
A =
µmuH
kB ln 10
B = log10
µmuP
kB
Article number, page 18 of 21
(B.3)
(B.4)
100101r (R∗)10-710-610-5s (m)CCCC100101r (R∗)10-710-610-5s (m)gSigSigSigSi1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)20.022.525.027.530.032.535.037.5Cumulative time [log(yrs)]É. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
ferential size distribution proportional to s−3.5, and summed up
to give a complete map of cumulative times, covering all grain
sizes. This is shown in figure C.3 and can be directly compared
to the map obtained with the dynamical code (Fig. C.1).
We see that the 2D (r,s) map obtained with the dynamical
code better follows the impact of changing β values due to subli-
mation, especifically it better captures the dynamics of the grains
just above the blow-out size. These grains produce the so-called
(and well-documented) pile-up close to the sublimation distance,
that is however not recover in the 2D map built using typical
timescales. The two cumulative time maps are then combined to
produce a single, smooth map, where averaged values are taken
when the bins are non-zero in both maps. This combined map
can be regarded as a density map.
To get an optical depth map, the density map is multiplied
by the geometrical cross-section in each bin, and divided by the
distance to the star. This optical depth is thus vertically and az-
imuthally integrated. This can be used to construct the optical
depth radial profile by integrating over all grain sizes for exam-
ple, or as a pre-requisite to evaluate the flux in scattered light and
thermal emission (e.g. Figure 8). The 2D maps can also be used
to truncate the disk, to estimate the emission within a certain
radius for example.
For the sake of comparison, in all the paper the maps shown
are normalized to get a dust disk to star flux ratio of 1% at λ =
2 µm.
Appendix D: Flux level produced by an evaporating
exocomet
Here, we describe the methodology employed to compute the
amount of dust released by the exocomets and the resulting flux
density. The successive steps are the following :
1. for each set of exocometary orbital parameters (see Sec. 4.1
and Tab. 4), we calculate the total mass of dust per unit of
exocomet surface (in kg/m2) released by the exocomet in one
complete revolution. This is done by integrating equation 17
of Marboeuf et al. (2016) along the exocomet orbit over one
orbital period. The exocomet is supposed composed at 50%
of dust. The results are independent of the actual size of the
exocomet, but depend on the assumed grain size distribution.
In the model of Marboeuf et al. (2016), the mass is calculated
assuming a differential grain size distribution proportional to
s−3.5 between smin = 1 µm and smax = 1 mm.
2. an exocomet radius is assumed to get the exocomet surface
and hence the total mass of dust ejected from the exocomet
in one orbit using the result of the previous step. The released
dust masses are documented in the 6th row of Tables D.1 and
D.2, assuming an initial exocomet radius of 10 km.
3. we make the approximation that all the dust ejected by the
exocomet during one orbit is released at perihelion. Although
this may appear a crude approximation, Fig. 10 shows that
this remains reasonable because a large fraction of the mass
is produced close to the star, due to the high eccentricities
considered in our study, and because the mass loss rate de-
creases as the distance squared to the star in the innermost re-
gions (inside the water ice sublimation distance in the model
of Marboeuf et al. 2016, see their Eq. 17).
4. this mass is distributed over the grain size bins of our model
grid (see App. C) between 1 mm downto the smallest grain
size in our simulation, assuming a differential grain size dis-
tribution proportional to s−3.5. This step allows each size bin
to be populated with an absolute number of grains. Note that
Article number, page 19 of 21
Fig. C.2: Dominant physical processes driving the evolution of
a grain, as a function of the distance to the star and grain size,
with red being inward migration by PR-drag, green being subli-
mation, orange being a combination of both and blue being ejec-
tion by radiation pressure (see text for more detail on how the
dominant process is estimated).
Fig. C.3: Cumulative time map derived from the estimated
timescales for the A0, carbon case.
PR-drag migration timescales are comparable. This is illustrated
in figure C.2, where we can see that sublimation is the domi-
nant physical process close to the sublimation distance for the
smallest grains, while inward migration is the dominant physical
process for the biggest grains, and ejection dominates over the
other processes otherwise.
This second 2D map provides crude evolution tracks for the
grains that are used to populate another cumulative time map
similar to the one shown in figure C.1. We proceed as follows.
For each size bin, we populate the initial position bin given the
value of β, thereby accounting for the eccentricity of the orbit
while the apoastron is fixed by the parent belt position. Then,
we follow each synthetic grain along the (r, s) plane, with a path
determined by the smallest timescale in each of the successive
bins through which the grain is passing. Each local, smallest
timescale is stored after being multiplied by the initial dust dif-
1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)1e-021e-011e+001e+011e+02r (au)1e-091e-081e-071e-061e-051e-041e-03Grain size (m)22.525.027.530.032.535.037.540.0Estimated cumulative time [log(yrs)]A&A proofs: manuscript no. V_6.1
the smallest size of our grid is lower than the smin = 1 µm
lower limit adopted by Marboeuf et al. (2016), but the frac-
tion of the total mass contained in the smallest grains is neg-
ligible with the adopted size distribution. Moreover, only the
bound grains are kept in the next steps.
5. this setup is used to produce a 2D map of cumulative times as
described in Appendix C, weighted by the absolute number
of grains of each size obtained at the previous step. The 2D
map is obtained by evolving the grains in position and size
over the largest lifetime of the biggest grains (3rd row of
Tables D.1 and D.2), see Appendix C for the methodology.
This step assumes in essence a system at steady state and is
equivalent to populating the orbits to produce a density map.
The 2D map is normalized by the time over which the simu-
lation was evolved (3rd row of Tables D.1 and D.2) to obtain
a mean 2D number density map equivalent to a density map
assuming a constant dust production process at perihelion.
This procedure is valid if the grains released at the comet's
perihelion which dominate the flux survive long enough for
their positions to be randomized in longitude along their or-
bits. We have checked that, in the case of the A0 star and
carbon grains, the orbital periods vary by a factor five within
the considered grain range. This guaranties randomization in
longitude within a few orbits, i.e. ∼1000 years in this case,
to be compared with the grain lifetime which is typically two
orders of magnitude larger.
6. the mean 2D number density map is used to directly com-
pute the absolute scattered light and thermal emission at
the wavelength of 2 µm, providing an estimate of the mean
flux density resulting from the first passage of a exocomet
(7th row in Tables D.1 and D.2). As time goes one, the exo-
comet radius shrinks. The number of exocomet orbits before
complete sublimation and the exocomet lifetimes are docu-
mented in the 5th and 4th rows of Tables D.1 and D.2, re-
spectively. The total flux density at 2 µm, produced by all the
grains released by the exocomet over its lifetime, is then ob-
tained by summing up the contributions at each successive
perihelion passage until complete erosion of the exocomet
(8th row of Tables D.1 and D.2). The flux at 10 µm (9th row
of Tables D.1 and D.2) is computed in a similar way.
Article number, page 20 of 21
É. Sezestre et al.: Hot exozodiacal dust : an exocometary origin?
Table D.1: Parameters and flux for the A0 and F0 stars, resulting from the exocomet evaporation model, assuming a 10 km-sized
exocomet. See Appendix D for details.
Star
Composition
Maximum grain lifetime (yr)
Exocomet lifetime (yr)
Number of exocometary orbits
Mass released in one orbit (kg)
Flux ratio at 2 µm (first orbit)
Total flux ratio at 2 µm
Total flux ratio at 10 µm
Star
Composition
Maximum grain lifetime (yr)
Exocomet lifetime (yr)
Number of exocometary orbits
Mass released in one orbit (kg)
Flux ratio at 2 µm (first orbit)
Total flux ratio at 2 µm
Total flux ratio at 10 µm
48
Carbon
6.3 × 104
3.5 × 103
1.5 × 1014
2.0 × 10−6
3.3 × 10−5
8.4 × 10−4
A0
Astrosilicate Glassy silicate
1.8 × 105
8.6 × 103
6.0 × 1013
1.2 × 10−6
4.8 × 10−5
1.0 × 10−2
5.5 × 105
6.9 × 103
7.4 × 1013
3.6 × 10−7
1.1 × 10−5
2.6 × 10−3
117
95
204
Carbon
9.9 × 104
2.0 × 104
3.5 × 1013
3.1 × 10−6
2.1 × 10−4
4.0 × 10−3
Table D.2: Same as Tab. D.1 for the G0 and K0 stars.
F0
Astrosilicate Glassy silicate
1.4 × 105
3.5 × 104
2.0 × 1013
2.2 × 10−6
2.7 × 10−4
3.4 × 10−2
2.1 × 105
1.7 × 104
4.4 × 1013
4.2 × 10−9
2.2 × 10−7
1.6 × 10−5
169
357
351
Carbon
1.8 × 105
4.3 × 104
2.5 × 1013
2.0 × 10−6
1.8 × 10−4
3.1 × 10−3
G0
Astrosilicate Glassy silicate
2.0 × 105
7.3 × 104
1.2 × 1013
1.9 × 10−6
3.8 × 10−4
3.7 × 10−2
3.3 × 104
3.0 × 104
5.4 × 1013
2.1 × 10−10
9.2 × 10−9
4.1 × 10−7
598
243
683
Carbon
1.8 × 105
9.6 × 104
2.6 × 1013
5.4 × 10−6
4.8 × 10−4
7.2 × 10−3
K0
Astrosilicate Glassy silicate
1.7 × 105
2.0 × 105
1408
5.5 × 1012
3.4 × 10−6
1.5 × 10−3
9.7 × 10−2
6.6 × 105
7.0 × 104
5.1 × 1013
2.8 × 10−10
1.3 × 10−8
2.5 × 10−7
496
Article number, page 21 of 21
|
1508.00427 | 1 | 1508 | 2015-08-03T14:19:14 | A statistical search for a population of Exo-Trojans in the Kepler dataset | [
"astro-ph.EP"
] | Trojans are small bodies in planetary Lagrangian points. In our solar system, Jupiter has the largest number of such companions. Their existence is assumed for exoplanetary systems as well, but none has been found so far. We present an analysis by super-stacking $\sim4\times10^4$ Kepler planets with a total of $\sim9\times10^5$ transits, searching for an average trojan transit dip. Our result gives an upper limit to the average Trojan transiting area (per planet) corresponding to one body of radius $<460$km at $2\sigma$ confidence. We find a significant Trojan-like signal in a sub-sample for planets with more (or larger) Trojans for periods $>$60 days. Our tentative results can and should be checked with improved data from future missions like PLATO2.0, and can guide planetary formation theories. | astro-ph.EP | astro-ph |
Draft version September 13, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
A STATISTICAL SEARCH FOR A POPULATION OF EXO-TROJANS IN THE KEPLER DATASET
Michael Hippke
Luiter Strasse 21b, 47506 Neukirchen-Vluyn, Germany
NASA Postdoctoral Program Fellow, NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
Daniel Angerhausen
Draft version September 13, 2018
ABSTRACT
Trojans are small bodies in planetary Lagrangian points. In our solar system, Jupiter has the largest
number of such companions. Their existence is assumed for exoplanetary systems as well, but none
has been found so far. We present an analysis by super-stacking ∼ 4× 104 Kepler planets with a total
of ∼ 9 × 105 transits, searching for an average trojan transit dip. Our result gives an upper limit to
the average Trojan transiting area (per planet) corresponding to one body of radius < 460km at 2σ
confidence. We find a significant Trojan-like signal in a sub-sample for planets with more (or larger)
Trojans for periods >60 days. Our tentative results can and should be checked with improved data
from future missions like PLATO 2.0, and can guide planetary formation theories.
the University Observatory Heidelberg discovered a new
"planet" 55◦ east of Jupiter and immediately noted its
strange orbit: "the small change in R.A. is remarkable"1.
More such bodies were found in the same year, and it was
realized quickly that these are trapped in Jupiter's La-
grangian points. To distinguish them from the main belt
asteroids, which usually receive female names, it was de-
cided to name them after Greek heroes of the Trojan war.
Wolf's "planet" is today known as (588) Achilles, and is
in the L4 group.
Asteroids that are trapped in L4 or L5 orbit around
their point of equilibrium in a tadpole, or horseshoe or-
bit (Marzari et al. 2002). Today, > 6, 000 Jupiter Tro-
jans are known2, as well as a few Neptune-, Mars- and
Earth Trojans. The largest known Trojans have sizes
> 100km in radius (Fern´andez et al.
2003), and it is
believed that the total number of L4 Jupiter Trojans,
with radii > 1km, is ∼ 6 × 105 (Yoshida & Nakamura
2005). If L5 contains an equal amount of debris, then the
total transiting area equivalent of small Jupiter Trojans
is corresponding to one body of radius ∼600km. The 32
2003) account for an
largest objects (Fern´andez et al.
additional radius equivalent of ∼300km.
The Lagrangian points are stable over Gyr timescales,
as long as the planet is < 4% of the system mass (Murray
& Holman 1999). Most of the system mass is usually
concentrated in the host star, e.g. 4% of M(cid:12) is 40MJup,
so that this limit is usually met. Consequently, we might
assume that other planetary systems also posses Trojan
bodies; this is also expected from formation mechanisms
in protoplanetary accretion discs (Laughlin & Chambers
2002). As the properties of extrasolar systems are di-
verse, we can ask the question of how large these bod-
ies can be, and how many there are. There is nothing
that physically prevents them from occurring in larger
1 In original German language: "Bemerkenswerst ist (...) die
kleine RA.-Bewegung von TG"
2 IAU Minor Planet Center, http://www.minorplanetcenter.
list retrieved on 26-Apr
net/iau/lists/JupiterTrojans.html,
2015
Fig. 1. -- Lagrangian points L4 and L5 are at 60◦ from the planet.
1.
INTRODUCTION
Back in 1771, the mathematician Lagrange found a
solution of the three-body-problem for a primary planet
and an asteroid of small mass. When the bodies are in
the same plane in circular orbits of the same period, the
stable locations for the asteroid are 60◦ from the planet
(Lagrange 1772). As no such asteroids where known
at the time, the problem was considered to be only of
mathematical interest. Today, we refer to these points
as the (stable) Lagrangian points L4 and L5 (Figure 1,
based on Cornish (1998)).
More than a century later, Max Wolf
(1906) of
[email protected]
[email protected]
L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L5L1L2L3L4L52
Michael Hippke & Daniel Angerhausen
numbers (and/or larger sizes) than in our own system.
Hypothetical exo-Trojans have been shown to be stable
for up to Jupiter mass in the most extreme cases (´Erdi
et al. 2007), assuming low eccentricity (Dvorak et al.
2004).
Searching for Trojans in time-series photometry is dif-
ficult, as these bodies librate around their equilibrium
points to a substantial degree. This produces large tran-
sit timing variations, so that they are missed in standard
planet-search algorithms. Also, the mean inclination of
Jupiter Trojans is 10◦ (Yoshida & Nakamura 2005) to
14◦ (Jewitt et al. 2000). If this is typical for exo-Trojans,
only part of the swarm would go into transit.
Data from the Kepler space telescope have been
searched for individual Trojans, with a null result and
sensitivity down to ∼ 1R⊕ (Janson 2013). Another
search was carried out with data from the MOST satel-
lite for the transiting Hot Jupiter HD 209458b, also with
a null result and an upper limit of ∼ 1 lunar mass of
asteroids (Moldovan et al. 2010).
Although interesting, we do not repeat these searches
for individual Trojans here, but ask the question of the
average Trojan effect in all Kepler data. Millions of small
Trojans might show up, on average, when stacking ∼
4 × 104 planets with a total of ∼ 9 × 105 transits, as is
the case for exo-moons (Hippke 2015).
2. METHOD
We employed the largest database available: High pre-
cision time-series photometry from the Kepler space-
craft, covering 4 years of observations (Caldwell at al.
2010).
2.1. Data selection
Based on a list of all validated transiting Kepler plan-
ets (821) and un-validated planet candidates (3,359)
2011)3, we downloaded their Kepler
(Wright et al.
long-cadence (LC, 30min) datasets. We used the same
dataset as published by the Transiting Planet Search
(TPS) pipeline, which relies on a systematic error-
corrected flux time series from a "wavelet-based, adap-
tive matched filter that characterizes the power spectral
density (PSD) of the background process yielding the
observed light curve and uses this time-variable PSD es-
timate to realize a pre-whitening filter and whiten the
light curve"(Borucki et al.
2011). This dataset was
used for most planet validations (e.g., Lissauer et al.
(2012); Rowe et al. (2014)). We have downloaded these
data from the NASA Exoplanet Archive4 and applied no
further corrections or detrending.
It must be assumed
that there are unidentified transits, and stellar trends, in
these data, but we can also assume that these are dis-
tributed randomly over phase time, so that no systematic
effect should affect the precise locations of the L4 and L5
phase time.
2.2. Data processing
Each planet has its own light curve in this dataset,
which comes with companion transits removed (in mul-
tiple systems). We phase-folded every lightcurve with its
3 www.http://exoplanets.org, list retrieved on 18-Nov 2014
4 http://exoplanetarchive.ipac.caltech.edu/docs/API_tce_
columns.html, retrieved on 21-Apr 2015
published period. Afterwards, we re-normalized the data
for each curve, while masking the times around plane-
tary primary and secondary transit. Then, we re-binned
each phase-folded lightcurve in 1,000 bins. Depending
on the period, this is equivalent to a time of 1min (for
the shortest period) to 18hrs (for a 750-day period). For
the median period of 13 days, the bin length is 20min.
As the average transit duration is a few hours, smearing
occurs only for the few very long period planets.
2.3. The super-stack
sample
of
From the
3739 useful phase-folded
lightcurves in 1,000 bins, we created a super-stack by co-
adding these, and taking the median of each bin. This
method was also used by Sheets & Deming (2014) for the
detection of average secondary eclipses, and by Hippke
(2015) for the search of an average exo-moon effect. In
contrast to these studies, we did not stretch to the ex-
pected transit duration, because Trojans are expected to
be in orbits around the Lagrange-points, and not station-
ary in phase time.
The resulting data were strongly dominated by out-
liers. This is caused by many factors contributing to
different noise levels: The brightness of the host star,
stellar variability, instrumental differences, and others.
We decided on two filters to remove outliers, namely the
stellar brightness (we kept stars brighter than 15mag in
J as measured by 2MASS), and the scatter per star (we
kept the better half).
It is interesting to mention a (slight) selection effect
from this filter choice: When rejecting dimmer and/or
noisier stars, the average stellar radius changes. Smaller
stars (e.g. M-dwarfs) exhibit more stellar noise (Basri et
al. 2013) and are usually less luminous. Consequently,
our sample is shifted towards larger stellar radii. While
the total Kepler -planet sample has an average stellar ra-
dius of 1.14R(cid:12), our post-filter sample has an average of
1.17R(cid:12).
3. RESULTS
The initial post-filtered superstack does not show any
significant Trojan dips, as shown in Figure 2. When
taking the average flux in a bin of 0.03 width in phase
space, we obtain +0.06 ± 0.23ppm for L4, and −0.10 ±
0.23ppm for L5. For the average stellar radius of 1.17R(cid:12),
we can set an upper limit for the average Trojan area (per
planet) of 460km at 2σ confidence. This applies to the
full (filtered) Kepler sample.
3.1. Cross-check for secondary eclipses
As a useful cross-check for our data preparation
method, we have searched for the average secondary
eclipse. For simplicity, we have assumed only reflected
light with an average albedo of 0.22 (Sheets & Deming
2014) and neglected differences in temperatures. As
can be seen in Figure 2, there is only a hint of a sec-
ondary feature at phase 1.0 = 0.0, which is measured to
be −0.31±0.21ppm at a bin width of 0.01. Following our
simplified assumptions, we can calculate the expected dip
for this sample as (RP /a)2 per planet, giving an average
of -0.88ppm for the sample. We explain the difference as
caused by smearing from shifts in transit timing (from
non-zero eccentricity) and different transit durations of
each planet, which we did not compensate for.
Exo-Trojans
3
autocorrelation (p = 0.01) if the complete dataset of
1,000 bins is used. However when we excise the phase
times with signals (around 0, 0.33, 0.5, 0.66 and 1) and
treat only the remaining data, then autocorrelation is
insignificant even at the 10% level.
We have also cross-checked whether this dip is intro-
duced by some symmetry artifact. When selecting any
other phase-folded time, e.g. flux < 0 at phase 0.2, no
equivalent dip on the "other side" of the orbit, i.e. at
phase 0.8, can be reproduced. Figure 4 shows this: If an
artifact were present, we would expect a dip centered at
each red mark, which is not present. Clearly, we cannot
produce a similar dip at any phase time with this sym-
metry argument; it only works at phase times 0.33 (L4)
and 0.66 (L5). We caution, however, that the S/N of the
total signal is low, as will be explained in the following
section. Splitting such a weak signal into different views
can therefore only create weak indicators of its validity.
3.3. Significance of the result
To measure the significance of these dips, we take the
2002;
2014), which compares the depth of the
signal-to-noise ratio for transits (Jenkins et al.
Rowe et al.
transit mode compared to the out-of-transit noise:
(cid:112)
S/N =
NT
Tdep
σOT
(1)
with NT as the number of transit observations, Tdep
the transit depth and σOT the standard deviation of out-
of-transit observations. For the Lagrangian signals, we
find the L4 and L5 dip at S/N∼6.7 each, and a com-
bined S/N=9.3. It has been argued by Fressin et al.
(2013) that the detection of transits becomes unreliable
for a S/N (cid:46) 10, so that this signal can only qualify as a
tentative detection.
3.4. Sub-sample properties
We have compared the properties of the 1,940 planets
in our sub-sample to the total Kepler sample. We use
a nonparametric density estimation with a local polyno-
mial regression to include local confidence bands (e.g.,
Ruppert et al.
(2003); Takezawa (2006)).
We find a correlation of the Trojan-like dips to the pe-
riod of the host star: At p > 20 days, the probability
density moves towards more pronounced Trojan-like sig-
nal, but the effect becomes only formally significant for
60 < p < 350 days. This might either reflect a stability
dependence of Trojan bodies to their semi-major axis,
or a formation bias, or a mix of both. Due to radia-
tion effects (such as the Yarkovsky effect (Bottke et al.
2006), which can cause small objects to undergo orbital
changes), we can expect few (if any) close-in (p < 10
days) small asteroids. For long periods, p > 350 days,
the sample size is too small for a significant result.
We have tried the same density estimates for several
other measures, but all are formally insignificant. At
first glance, for example, one might assume that the im-
pact parameter of the planetary transit could positively
affect trojan detectability: For more central planetary
transits, one might assume a sky-coplanar trojan to also
(more likely) transit centrally, making detection easier.
This is, however, likely overpowered by the inclination
Fig. 2. -- The initial superstack shows no significant dips at the
Lagrangian points.
We have also checked a different sample which is ex-
pected to yield a higher secondary dip: All planets with
radii > 2R⊕ on orbits < 0.3au. This sub-sample is ex-
pected to yield an average secondary eclipse of -1.7ppm;
our data analysis gives −0.73 ± 0.34 in the same bin
width. Again, we have to expect centering variations
which reduce (broaden) the observed depth. It is how-
ever reassuring to see a dip at > 2σ significance. Fi-
nally, we have checked the few individual examples from
Sheets & Deming (2014) where the secondary eclipse is
detected for individual planets (e.g. Kepler-10b); these
dips are also present in our dataset. We conclude that
there seems to be no obvious fault in our dataset, and
that secondary eclipses are hard to detect for most of the
Kepler planets.
3.2. Sub-sample analysis
The full sample might be heavily diluted by a large
number of systems with no, or relatively few, Trojans.
We test this hypothesis by assuming that the flux in L4
and L5 is uncorrelated for any other cause than Trojan
bodies. We know from our own solar system that the
number of bodies in L4 and L5 is approximately equal.
Then, we can examine a sub-sample of the superstack:
We take all those planets that exhibit a negative flux at
L4 (phase 0.33), and take their data of phase 0.5..1 for
further analysis. The same is done for L5 in the reverse
logic. This gives us 1, 251 samples of negative flux at
phase 0.33, and their light curves for the "right part",
i.e. flux phase 0.5..1. We also find 1, 266 samples with a
dip at phase 0.66, and take their part of the light curves
from phase 0..0.5. Afterwards, we stitch these halves
together, and obtain 1,940 lightcurves (some have dips in
both halves). The result of this sub-sample is shown in
Figure 3 and exhibits a clear dip at both L4 and L5, with
a maximum depth of 2ppm (970km radius equivalent). It
is re-assuring to see that this dip is not uniform; as can be
seen in the double-phase fold (right part of this figure),
its shape is elongated away from planetary transit, as
is expected for distributions from horseshoe and tadpole
orbits.
An alternative interpretation of the dips in figure 3
(left) would be numerical fluctuations, caused by auto-
correlation. Indeed, a Durbin-Watson test returns clear
L4L5-10-8-6-4-2024681000.10.20.30.40.50.60.70.80.91Relative flux (ppm)Phase4
Michael Hippke & Daniel Angerhausen
Fig. 3. -- Sub-sample superstack in normal (left) and double symmetric (right) phase fold, with expected orbit size shown for reference.
Note different vertical axes. Gray dots are 1,000 bins over phase space, black dots with error bars (right) are 100 bins for better visibility.
Fig. 4. -- Cross-check of sub-sample selection artifacts. In each
line, we select those data that have a dip on one side of phase
space, and plot their flux only for the other half of phase space.
For example, in the first row, all data are shown that have a dip at
phase 0.034 (at boxcar width 0.03). If an artifact were present, we
would expect a corresponding dip (red color) at phase 1 − 0.34 =
0.966, which is not seen. Also, we would expect such an artifact
to occur in every line, centered at the black marked, which is not
the case. Instead, we mainly see red dips occur at phase ∼ 0.66,
where the Trojan transits are expected. A few columns (0.025 in
phase time) around primary transits are excised as these values
are ∼ 1000× deeper in flux, making them incompatible with useful
color scalings.
scatter of possible trojan "swarms". If we take the incli-
nation scatter of Jupiter's trojans as a proxy, then the
projected size of a trojan cloud in exoplanetary systems
would be several times larger than the projected size of
their star. Given our data quality, it is not surprising to
find no significant correlation with respect to the impact
parameter.
The same argument can be made for a correlation to
the stellar radius. One might hypothesize that the total
trojan mass correlates to the total mass of the circumstel-
lar disk, which itself might be correlated with the mass
of the star. More massive stars are known to host more
massive planets (Johnson et al.
2010), so that Trojan
bodies might also be more massive (i.e., larger at a given
Fig. 5. -- Density estimate for the Trojan-like signal versus plan-
etary period. The shaded area is the local 2σ confidence band.
Short periods < 50 days are consistent with zero trojan signal, but
a trojan-like signal is detected for periods between 60 and 350 days.
Uncertainty increases for longer periods due to a lack of data. See
text for discussion.
density). The scale for such a correlation, however, is
expected to be of order Rtrojan ∼ M 1/3
trojan. The major-
ity of Kepler planets and candidates are found for star
between 0.5R∗ and 2R∗, a range which, in combination
with the limited data quality, does not allow for the de-
tection of a significant correlation of trojan occurrence
to the stellar radius.
Finally, we also tried correlations to the planetary ra-
dius, multiplicity, metallicity of the host star, and eccen-
tricity; all of which gave null results.
4. DISCUSSION
While the data quality from Kepler is the best we have,
it is only barely sufficient to search for Trojan bodies.
Still, we believe that the methods outlined in this paper
will be valuable in the future, when more and better data
come available. The PLATO 2.0 mission (Rauer et al.
2014) will deliver photometry for 500 bn stars in the years
after 2024, and up to 3× better photometric precision.
With such a dataset, the analysis performed here should
be repeated, and should yield highly significant results
L4PrimarytransitL5-20-15-10-50510152000.10.20.30.40.50.60.70.80.91Relative flux [ppm]PhaseL4 L5 -5-4-3-2-10123450.50.60.70.80.91Relative flux [ppm] Phase Phase0.510.90.60.80.7-20-15-10-5051015200100200300400Lagrangian flux [ppm]Planetary period [days]Exo-Trojans
5
for every breakdown. Also, a few single large Trojans
might also be expected, if the vast dataset (Hippke &
Angerhausen 2015) can be mined sufficiently.
We have explored the potential of PLATO 2.0 using
(2014) for its
lower-limit estimates from Rauer et al.
scientific return. Then, ∼ 10× more lightcurves will be
available, when compared to Kepler, for a duration of 6
(instead of 3) years. For simplicity, we neglect the bet-
2 × 10
ter instrumental noise properties. This gives 1/
of noise improvement per bin (compared to the current
data), resulting in ∼ 0.3ppm of noise in each of the 1,000
bins. Consequently, from a superstack without any se-
lections, we can expect equal or better signal-to-noise
properties than in the heavily selected and biased Kepler
sample. More precisely, the full PLATO 2.0 sample is
expected to yield a signal-to-noise as shown in Figure 3,
without any of our discerning selection choices.
√
2014).
Furthermore, we can expect to make clear detections of
large individual Trojans with PLATO 2.0, if such bodies
exist with transiting areas > 0.5R⊕. To show this, we
have created a series of injections. Our process is similar
to the one described in Hippke & Angerhausen (2015).
In short, we take real solar data from VIRGO/DIARAD
(Frohlich et al. 1997) and add instrumental noise from
an end-to-end PLATO 2.0 simulator (Zima et al. 2006;
Marcos-Arenal et al.
Into these data, we in-
ject synthetic Trojan lightcurves, assumed to orbit in
horseshoe-orbits around L4/L5 with semi-major axes of
∼ 0.02 in phase-time (Janson 2013). To explore the pa-
rameter space of recoverable signals, we varied the tran-
siting Trojan area (radius), the stellar radius, and the
planetary period. We show an exemplary riverplot in
Figure 6 for a Sun-like star, orbited by a 10-day Hot
Jupiter and a 0.8R⊕ Trojan in L4. Such a configura-
tion is easily detected in visual examination, but might
escape standard transit detection algorithms due to the
large transit timing variations. We find that for Sun-like
stars, transiting areas of > 0.65R⊕ (Mars-size) are easily
detected visually; the equivalent limit for a 0.5R(cid:12) M-
Dwarf is ∼0.5R⊕. Instead of a visual search, algorithms
might be used, as explained by Janson (2013). It is how-
ever unclear how efficient these can be; this would have
to be determined with a series of injections and (blind)
retrievals.
An interesting finding from our own injections is that
potential Trojans at shorter period planets are much
more easily identified due to their higher number of tran-
sits. Our example uses a 10-day planet; this is on the
lower end of our hypothesized limit of planets having
Trojans, as explained in section 3.2. For longer period
planets, e.g. a 100-day planet, the number of rows (tran-
sits) in Figure 6 would be 20 (instead of 200) for 6 years
of data. This reduces chances of detection, highlighting
the benefits of long-term campaigns.
5. CONCLUSION
With the given dataset, we only find a significant
Trojan-like signal when applying the "left-right" method,
selecting only L5 data for those candidates that seem to
exhibit a L4 dip (and vice versa). While we tested this
method to be robust against a symmetrical bias, it also
implies that the main sample must be heavily diluted
with a large number of systems with no (transiting) Tro-
jans. If the method is valid, then the breakdowns of this
Fig. 6. -- Riverplot of Trojan injection into 6 years of real so-
lar data, assuming PLATO 2.0 instrumental performance. A syn-
thetic 10-day period Hot Jupiter is visible at phase 0.5; the injected
0.8R⊕ Trojan is clearly visible in its horseshoe-orbit at L4.
sub-sample indicate that Trojans are more prominently
found for longer (>60 days) periods. These cautious, and
preliminary findings might inspire theorists to advance
planetary formation theory in that direction; these theo-
ries can then be validated with the upcoming data from
the PLATO 2.0 spacecraft.
REFERENCES
Basri, G., Walkowicz, L. M., Reiners, A. 2013, ApJ, 769, 1
Borucki, W., Koch, D. G., Basri, G. 2011, ApJ, 736, 19
Bottke, W. F., Vokrouhlick´y, D., Rubincam, D., et al. 2006,
Annual Review of Earth and Planetary Sciences, 34,
Caldwell, D. A., Kolodziejczak, J. J., Van Cleve, J. E., et al.
Janson, M. 2013, ApJ, 774, 2
Jenkins, J.M., Caldwell, D.A., Borucki, W. J. 2002, ApJ, 564, 495
Jewitt, D. C., Trujillo, C. A., Luu, J. X. 2000, A&A, 120, 2
Johnson, J. A., Aller, K. M., Howard, A. W. 2010, PASP, 122, 894
Kolmogorov, A. 1933, Giornale dell'Istituto Italiano degli Attuari,
2010, ApJ, 713, L92-L96
Cornish, N. J. 1998, online:
http://map.gsfc.nasa.gov/ContentMedia/lagrange.pdf
´Erdi, B., Nagy, I., S´andor, Z., et al. 2007, MNRAS, 381, 1
Dvorak, R., Pilar-Lohinger, E., Schwarz, R., et al. 2004, A&A,
426, L37
Fern´andez, Y. R., Sheppard, S. S., Jewitt, D. C. 2003, AJ, 126, 3
Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, 2
Frohlich, C., Andersen, B. N., Appourchaux, T. 1997, Sol. Phys.,
170, 1
Hippke, M. 2015, ApJ, in press, arXiv:1502.05033
Hippke, M., Angerhausen, D. 2015, ApJ, submitted,
arXiv:1503.03251
4, 8391
Lagrange, J.-L. 1772, Essai sur le probl`eme des trois corps, in:
Oeuvres de Lagrange, Gauthier-Villars, 229334
Laughlin, G., Chambers, J. E. 2002, AJ, 124, 1
Lissauer, J. J., Marcy, G. W., Rowe, J. F. 2012, ApJ, 750, 2
Marcos-Arenal, P., Zima, W., De Ridder, J., et al. 2014, A&A,
566, A92, 12
Marzari, F., Scholl, H., Murray, C, et al. 2002, Origin and
Evolution of Trojan Asteroids, W. F. Bottke (ed.), University
of Arizona Press, Tucson, 725-738
Moldovan, R., Matthews, J. M., Gladman, B., et al. 2010, ApJ,
716, 1
Mulders, G.D., Pascucci, I., Apai, D. 2015, ApJ, 798, 2
0PXδt200PXnPXPhase 0.33 0.5 0.66 (L4)6
Michael Hippke & Daniel Angerhausen
Murray, N., Holman, M. 1999, Science, 283, 5409
Rauer, H., Catala, C., Aerts, C. et al. 2014, Experimental
Astronomy, 38, 1-2
Takezawa, K. 2006, Allgemeines Statistisches Archiv, 90, 4
Wolf, M. 1996, Astronomische Nachrichten, 170, 353
Wright, J. T., Fakhouri, O., Marcy, G. W., et al. 2011, PASP,
Rowe, J. F., Bryson, S. T., Marcy, G. W., et al. 2014, ApJ, 784,
123, 902
1, 45
Ruppert, D. R., Wand, M. P., Carroll, R. J. 2003, Semiparametric
Regression, Cambridge University Press
Sheets, H. A., Deming, D. 2014, ApJ, 794, 2
Smirnov, N. 1948, Annals of Mathematical Statistics, 19, 279281
Yoshida, F., Nakamura, T. 2005, AJ, 130, 6
Zima, W., Arentoft, T., De Ridder J., et al. 2010, Astronomische
Nachrichten, 88, 789-793
|
1201.3029 | 1 | 1201 | 2012-01-14T17:09:14 | Multiband Optical Observation of P/2010 A2 Dust Tail | [
"astro-ph.EP"
] | An inner main-belt asteroid, P/2010 A2, was discovered on January 6th, 2010. Based on its orbital elements, it is considered that the asteroid belongs to the Flora collisional family, where S-type asteroids are common, whilst showing a comet-like dust tail. Although analysis of images taken by the Hubble Space Telescope and Rosetta spacecraft suggested that the dust tail resulted from a recent head-on collision between asteroids (Jewitt et al. 2010; Snodgrass et al. 2010), an alternative idea of ice sublimation was suggested based on the morphological fitting of ground-based images (Moreno et al. 2010). Here, we report a multiband observation of P/2010 A2 made on January 2010 with a 105 cm telescope at the Ishigakijima Astronomical Observatory. Three broadband filters, $g'$, $R_c$, and $I_c$, were employed for the observation. The unique multiband data reveals that the reflectance spectrum of the P/2010 A2 dust tail resembles that of an Sq-type asteroid or that of ordinary chondrites rather than that of an S-type asteroid. Due to the large error of the measurement, the reflectance spectrum also resembles the spectra of C-type asteroids, even though C-type asteroids are uncommon in the Flora family. The reflectances relative to the $g'$-band (470 nm) are 1.096$\pm$0.046 at the $R_c$-band (650 nm) and 1.131$\pm$0.061 at the $I_c$-band (800 nm). We hypothesize that the parent body of P/2010 A2 was originally S-type but was then shattered upon collision into scaterring fresh chondritic particles from the interior, thus forming the dust tail. | astro-ph.EP | astro-ph |
Multiband Optical Observation of P/2010 A2 Dust Tail
Junhan Kim1
3##-###, Mokdong Apartment, Mok-5-dong, Yangcheon-gu, Seoul 158-753, Republic of Korea
Masateru Ishiguro2
Department of Physics and Astronomy, Seoul National University, San 56-1, Silim-dong,
Gwanak-gu, Seoul 151-742, Republic of Korea
Hidekazu Hanayama
Ishigakijima Astronomical Observatory, National Astronomical Observatory of Japan, Ishigaki,
Okinawa 907-0024, Japan
Sunao Hasegawa and Fumihiko Usui
Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1
Yoshinodai, Chuo-ku, Sagamihara, Kanagawa 252-5210, Japan
Kenshi Yanagisawa
Okayama Astrophysical Observatory, National Astronomical Observatory of Japan, Asaguchi,
Okayama 719-0232, Japan
Yuki Sarugaku
Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1
Yoshinodai, Chuo-ku, Sagamihara, Kanagawa 252-5210, Japan
Jun-ichi Watanabe
National Astronomical Observatory of Japan, Mitaka, Tokyo, 181-8588, Japan
and
Michitoshi Yoshida
Hiroshima Astrophysical Science Center, Hiroshima University, Higashi-Hiroshima, Hiroshima
739-8526, Japan
ABSTRACT
An inner main-belt asteroid, P/2010 A2, was discovered on January 6th, 2010. Based
on its orbital elements, it is considered that the asteroid belongs to the Flora collisional
-- 2 --
family, where S-type asteroids are common, whilst showing a comet-like dust tail. Al-
though analysis of images taken by the Hubble Space Telescope and Rosetta spacecraft
suggested that the dust tail resulted from a recent head-on collision between asteroids
(Jewitt et al. 2010; Snodgrass et al. 2010), an alternative idea of ice sublimation was
suggested based on the morphological fitting of ground-based images (Moreno et al.
2010). Here, we report a multiband observation of P/2010 A2 made on January 2010
with a 105 cm telescope at the Ishigakijima Astronomical Observatory. Three broad-
band filters, g′, Rc, and Ic, were employed for the observation. The unique multiband
data reveals that the reflectance spectrum of the P/2010 A2 dust tail resembles that of
an Sq-type asteroid or that of ordinary chondrites rather than that of an S-type aster-
oid. Due to the large error of the measurement, the reflectance spectrum also resembles
the spectra of C-type asteroids, even though C-type asteroids are uncommon in the
Flora family. The reflectances relative to the g′-band (470 nm) are 1.096±0.046 at the
Rc-band (650 nm) and 1.131±0.061 at the Ic-band (800 nm). We hypothesize that the
parent body of P/2010 A2 was originally S-type but was then shattered upon collision
into scaterring fresh chondritic particles from the interior, thus forming the dust tail.
Subject headings: comets: general -- comets: individual(P/2010 A2) -- minor planets,
asteroids: general
1.
Introduction
Main-belt asteroid P/2010 A2 (LINEAR, hereafter P/2010 A2) was discovered on January
6th, 2010 (Birtwhistle et al. 2010) by Lincoln Near-Earth Asteroid Research (LINEAR). It showed
a comet-like dust tail without a central coma, representing a remarkable distinction from ordinary
comets (Moreno et al. 2010; Jewitt et al. 2010; Snodgrass et al. 2010; Hainaut et al. 2011). The
mechanism of the comet-like activity remains under debate. Moreno et al. (2010) applied inverse
dust tail Monte Carlo fitting to their observational images and argued that water-ice sublimation
brought about the cometary activity. However, three independent research groups stated that colli-
sional disruption was responsible for the dust tail on the basis of their observations and dynamical
studies (Jewitt et al. 2010; Snodgrass et al. 2010; Hainaut et al. 2011). Note that more than 10
months had already passed since the dust emission. Details such as impact ejecta plume and jet
structure may not have been captured in their observed images (Hsieh et al. 2011; Bodewits et al.
2011; Ishiguro et al. 2011).
The proper orbital elements of P/2010 A2 are a semi-major axis a = 2.29 AU, an eccentricity
e = 0.12, and an inclination i = 5.26◦. It belongs to a dynamical asteroid group, the Flora family,
1Graduated from Seoul National University, Republic of Korea in August 2010
2Corresponding author. E-mail : [email protected]
-- 3 --
whose proper orbital elements are 2.12 < a < 2.31AU, 0.11 < e < 0.175, and 3◦ < i < 7.5◦
(Nesvorn´y et al. 2002). It is important to note that P/2010 A2 remains within the snowline of the
Solar System. The orbital aspect differentiates it from known main-belt comets, which have an
aphelion greater than 2.5 AU (Hsieh & Jewitt 2006; Hsieh 2009; Jewitt 2011). In addition to their
dynamical properties, asteroids are classified in accordance with the shape of their spectra (e.g.,
Bus et al. 2002). Furthermore, asteroid families share similar spectral properties, and therefore,
family members belong to the similar taxonomic types (Cellino et al. 2002). Most of the Flora
members are categorized as S-type asteroids (Florczak et al. 1998). However, S-type asteroids are
known to be parent bodies of ordinary chondrites, not containing obvious evidence for water ice,
unlike C-type or D-type asteroids (Campins et al. 2010; Licandro et al. 2011).
To date, the taxonomic type of P/2010 A2 has not been reported in refereed journals. In this
paper, we present the photometric observation of P/2010 A2 made at the Ishigakijima Astronomical
Observatory, Japan, by simultaneous imaging with g′, Rc, and Ic-band filters. We describe the data
reduction and calibration process used to derive the reflectance spectrum of the object from the
integrated fluxes at different wavelengths. By comparing the reflectance to the spectra of different
types of known asteroids, we discuss the cause of the comet-like activity of P/2010 A2.
2. Observation and Data Reduction
All P/2010 A2 images used in this research were obtained on 15th January, 2010 (15:56 --
20:02 UT) with a 105 cm telescope of Ishigakijima Astronomical Observatory (IAO), National
Astronomical Observatory of Japan (NAOJ), located on Ishigaki Island, Okinawa, Japan. On the
observation date, the object's heliocentric distance was 2.014 AU, geocentric distance was 1.043
AU, and solar phase angle was 6.45◦. The telescope at IAO, named 'Murikabushi' has a 105 cm
diameter. The optical system is a Ritchey-Chr´etien and the focal length is 1260 cm (F12). At
the Cassegrain focal plane, a three-channel simultaneous imaging system with three CCD cameras
(MITSuME: Multicolor Imaging Telescopes for Survey and Monstrous Explosions) was attached
for the observation. This system has band-pass filters of SDSS (Sloan Digital Sky Survey) g′,
Johnson -- Cousins Rc, and Ic. Each CCD camera consists of an Alta U6 (Apogee Instruments Inc.)
with an array size of 1024 × 1024 pixels and a pixel size of 24 × 24 µm. We took 44 images for
each wavelength under clear conditions: 24 images were taken with an exposure time of 180 s, and
a further 20 images were attained with an exposure time of 300 s. In total, we have an integration
time of 172 minutes.
Concerning the data reduction, we focused on subtracting background stars from all the images
to obtain a faint dust tail (Ishiguro et al. 2007a; Ishiguro 2008). This was accomplished using our
own programs to remove or mask stars and also using well-known software such as IRAF, SExtractor
(Bertin & Arnouts 1996), and WCStools (Mink 1997). Firstly, we prepared star-aligned composite
images using WCS (World Coordinate System) information in order to detect background sources
including stars and galaxies. WCS information was necessary to precisely determine the locations
-- 4 --
of the object in every frame. Then, bright sources other than the object were automatically found
by SExtractor. A star-aligned image is more useful than a single image because it enables easier
detection of faint stars. Then, circular masks, with zero-values and sizes corresponding to the
width of the point spread function of individual stars, were applied to the location of stars (see
Figure 1). Finally, combining all the masked frames with offsets to align the images with respect
to the position of P/2010 A2 gave us an image showing only a faint dust tail. In order to calculate
the offsets, the position of the object was computed at a given time using our program, and the
sky2xy command in W CStools was utilized to convert the equatorial coordinate to a CCD image
coordinate. The masked images were combined with offsets to give the image of the dust tail,
excluding the masked pixels and shifting the background intensity to a zero level. Since P/2010
A2 moved relative to the background objects, it is possible to exclude the masked data completely,
and to combine the sequence of images. Finally, we obtained an image without stars and galaxies.
These resulting images are shown in Figure 2.
3. Analysis
3.1. Flux Calibration
To integrate a flux coming from the object for each wavelength, a cut profile along the perpen-
dicular direction of the dust tail was analyzed. Assuming an identical flux distribution for different
wavelength bands, a region was selected(see Figure 3(a)) to extract the profile. For precise measure-
ments on this extended structure, the flux was averaged along the line of the tail. This was done
using the projection function in on SAOImage ds9 (Joye & Mandel 2003). The background flux
coming from the sky brightness was fitted to a polynomial function and subsequently subtracted
from the original profile.
Figure 3(b) shows cut profiles for three wavelengths. Note that a raw intensity value (ADU,
arbitrary data unit) was used here to present the flux. This raw intensity value can be scaled to
other physical units using conversion parameters determined from standardization, as shown below.
3.2. Reflectance
In order to convert the ratio of flux for three wavelengths into relative reflectances, we need
photometric transformation equations as well as the color indices of the Sun. The transformation
equations were derived by photometry from standard star data. Unfortunately, our observation
of the standard field contained several stars whose standard magnitudes at g′-band are unknown.
Based on the magnitudes of standard stars listed in Landolt (1992), we found the UBVRcIc pho-
tometric magnitudes and converted them to Sloan u′g′r′i′z′ magnitudes. Conversion equations
between the two systems are given by Smith et al. (2002). After the arithmetic operations, a stan-
-- 5 --
dard star chart of the g′RcIc system was prepared for analysis, and then, the parameters for the
photometric transformation equations were obtained using the f itparams function inIRAF.
The solar color indices for various filter systems are well known. However, the filter system
used in our observation is a combination of two photometric systems : g′ of the SDSS system, Rc
and Ic of the Cousins system. Therefore, solar color indices such as (g′ − Rc)⊙ and (g′ − Ic)⊙ are
required. Again, transformation equations between the Johnson-Cousin (UBVRcIc) system and
the SDSS (u′g′r′i′z′) system of Smith et al. (2002) were used :
V = g′ − 0.55(g′ − r′) − 0.03
V − Rc = 0.59(g′ − r′) + 0.11
Rc − Ic = 1.00(r′ − i′) + 0.21
(1)
(2)
(3)
Combining the above equations and applying the solar indices of (g′ − r′)⊙ = 0.45 and (r′ −
i′)⊙ = 0.10 (Ivezic et al. 2001), we obtain (g′ − Rc)⊙ = 0.653 and (Rc − Ic)⊙ = 0.963
On the basis of these indices and transformation equations, we determined the relative re-
flectances compared to the reflectance at a particular wavelength (here, at g′-band : Ig′/I⊙,g′) as
follows:
I⊙,Rc(cid:19) /(cid:18) Ig′
(cid:18) IRc
I⊙,Ic(cid:19) /(cid:18) Ig′
(cid:18) IIc
I⊙,g′(cid:19) = 100.4×[(g′−Rc)−(g′−Rc)⊙]
I⊙,g′(cid:19) = 100.4×[(g′−Ic)−(g′−Ic)⊙]
(4)
(5)
Note that the characters shown on the left hand side of the equations denote intensities, which
are the integrated quantities of the flux profile in Figure 3(b). We had to exercise caution in
determining the reflectance, given as an exponential function with base 10, because it is highly
sensitive to color correction and transformation of the standard photometric system.
4. Results & Discussion
Table 1 shows the reflectance spectrum normalized at g′-band (470 nm). In order to deter-
mine the taxonomic type of the object, we calculated the average spectra of different taxonomic
type asteroids from the Flora family using SMASS II (Phase II Small Main-Belt Asteroid Spec-
troscopic Survey) data (Bus & Binzel 2002a). In the archive, there are 66 asteroids in the Flora
asteroid family whose spectral types are known. The members are selected according to the asteroid
dynamical family classifications (Zappal`a et al. 1995). Among them, we found that there are 61
S-type asteroids, including 4 Sq-type asteroids (1324 Knysna, 2873 Binzel, 2902 Westerlund, and
4733 ORO) and 2 C-type asteroids (2952 Lilliputia and 4396 Gressmann). In Figure 4(a) -- (c), the
-- 6 --
average reflectance spectra of S-type, C-type, and Sq-type asteroids with one standard deviation
are compared with that of P/2010 A2.
Firstly, we noticed that the reflectance spectrum of the P/2010 A2 dust tail is significantly
different from that of S-type asteroids (Figure 4(a)). In Figure 4(b), we compare the P/2010 A2
spectrum to the reflectances of C-type asteroids. A C-type asteroid has a relatively flat spectrum for
λ > 550 nm but exhibits absorption for λ < 400 nm. Although the P/2010 A2 spectrum is slightly
redder than that of C-type asteroids, we cannot rule out the possibility of it being a C-type asteroid
due to large uncertainties in our measurement. Figure 4(c) and Figure 4(d) show the reflectance
spectra of Sq-type (of SMASS II class) asteroids and that of ordinary chondrites, respectively. We
obtained the spectra of ordinary chondrites from Gaffey (2001). These are considered as analogs
of S-type asteroids but show differences in slope for λ < 700 nm and in depth of absorption
at λ ∼ 950 nm. S-type asteroids show a redder spectral slope at visible wavelength, and the
depth of the absorption band at ∼950 nm is shallower than that of ordinary chondrites and Sq-
asteroids(Chapman 2004; Sasaki et al. 2001; Ishiguro et al. 2007b). Ordinary chondritic particles
were also found on an S-type asteroid Itokawa(Noguchi et al. 2011). It is estimated that our result
concerning P/2010 A2 is similar to those of Sq-type asteroids and ordinary chondrites. In particular,
our spectrum has the closest spectral property to that of H5 chondrite.
The significance in our result is that the P/2010 A2 debris shows optical properties that are
different from those of S-type asteroids. It would have been reasonable to suppose that P/2010 A2
has a spectrum similar to that of S-type asteroids, as they are the dominant taxonomic asteroids
in the Flora family.
In fact, 92 % (61 out of 66 asteroids) of the Flora asteroids in SMASS II
are classified as either S-type or Sq-type (a subclass of S-type). Why is P/2010 A2 different from
S-type? As we mentioned above, there is a possibility of it being a carbonaceous asteroid, which
could contain hydrated minerals and possibly water ice. This seems reasonable as some of the
main-belt comets (MBCs), which were likely to be activated by sublimation of water and ice, show
similar spectra to those of C-type asteroids (Hsieh et al. 2009b).
Whereas we dismiss the possibility of P/2010 A2 being C-type, we suggest that its debris
particles have a composition similar to that of ordinary chondrites. It would be unreasonable to
assume that there was water ice in the warm region of the inner main-belt (Hainaut et al. 2011). As
the size of the parent body was small, with a diameter of 120 m (Jewitt et al. 2010; Snodgrass et al.
2010), icy particles should have been sublimated if had they existed inside the parent bodies.
In addition, dark primitive asteroids are rare in the vicinity of P/2010 A2. Furthermore, the
inconsistency with S-type asteroids seems to be feasible even though they are the major constituents
of the Flora family. If dust particles were generated by a catastrophic collisional process, a large
portion of the materials constituting the dust tail should originate from inside the parent body.
Such materials would be fresher than materials on the surface of S-type asteroids due to the effect
of space weathering. As pointed out by Jewitt et al. (2010); Snodgrass et al. (2010); Hainaut et al.
(2011), we support the hypothesis of a catastrophic collision. Our results of determining spectral
type of the dust tail may be weak in the sense that they are based on only three spectral points in
-- 7 --
the visible range that prevent a definite classification of the parent body or the debris composition.
Further observations at longer wavelength (around 0.9 -- 1.0 µm) are essential to distinguish Sq-type
asteroids and C-type asteroids.
5. Summary
A main-belt asteroid P/2010 A2 was discovered with an unusual comet-like dust tail. We
performed photometric observations of this object, using a 1 m telescope at the Ishigakijima As-
tronomical Observatory, to determine the reflectance properties in order to assign a spectral type.
1. For three different wavelengths, the optical images were combined after masking bright
sources other than the object. Integrated fluxes coming from the object on the composite image
gave the following reflectances relative to g′-band (470 nm): 1.096±0.046 at Rc-band (650 nm) and
1.131±0.061 at Ic-band (800 nm).
2. The proper orbital elements, namely, the semimajor axis, the eccentricity, and the inclina-
tion, led to the object being classified as part of the Flora asteroid family, where S-type asteroids are
common (estimated to account for 92% of the Flora asteroid family, from the SMASS II archive).
However, the reflectance spectrum of the P/2010 A2 dust tail is dissimilar to those of S-type
asteroids.
3. Comparing the reflectance spectrum of P/2010 A2 to those of three different taxonomic
types from the Flora asteroid family shows that there is a possibility of it being a C-type or an Sq-
type asteroid. In addition, the spectrum resembles that of ordinary chondrites, which are considered
as analogs of S-type asteroids.
4. We suggest that the dust tail of P/2010 A2 has a property similar to ordinary chondrites
and that the parent body could be an S-type asteroid.
If the dust particles were generated by
the collision, it may be considered that fresh materials from the parent body were scattered and
showed a similar spectra to that of ordinary chondrites that can be found on an S-type asteroid.
This research was supported by the National Research Foundation of Korea. S.H. is supported
by the Space Plasma Laboratory, ISAS, JAXA.
Bertin, E. & Arnouts, S. 1996, A&AS, 117, 393
REFERENCES
Birtwhistle, P., Ryan, W. H., Sato, H., Beshore, E. C., & Kadota, K. 2010, IAU Circ., 9105, 1
Bodewits, D., Kelley, M. S., Li, J.-Y., et al. 2011, ApJ, 733, L3
-- 8 --
Bus, S. J., & Binzel, R. P. 2002, Icarus, 158, 106
Bus, S. J., Vilas, F., & Barucci, M. A. 2002, Asteroids III, UArizona Press, pp. 169
Campins, H., Hargrove, K., Pinilla-Alonso, N., et al. 2010, Nature, 464, 1320
Chapman, C. R. 2004, Annual Review of Earth and Planetary Sciences, 32, 539
Cellino, A., Bus, S. J., Doressoundiram, A., & Lazzaro, D. 2002, Asteroids III, UArizona Press,
pp. 633
Florczak, M., Barucci, M.A., Doressoundiram, A., Lazzaro, D., Angeli, C. A., & Dotto, E. 1998,
Icarus, 133, 233
Gaffey, M. J. 2001, Meteorite Spectra. EAR-A-3-RDR-METEORITE-SPECTRA-V2.0. NASA
Planetary Data System.
Hainaut, O.R. et al. 2011, A&A(in press)
Hsieh, H. H. & Jewitt, D. 2006, Science, 312, 561
Hsieh, H. H. 2009, A&A, 505, 1297
Hsieh, H. H. & Jewitt, D. & Fernandez, Y. R. 2009, ApJ, 694, L111
Hsieh, H. H., Ishiguro, M., Lacerda, P., & Jewitt, D. 2011, AJ, 142, 29
Ishiguro, M., Sarugaku, Y., Ueno, M., Miura, N., Usui, F., Chun, M-Y., & Kwon, S. M. 2007,
Icarus, 189, 169
Ishiguro, M. et al. 2007, Meteoritics & Planetary Science, 42, 10, 1791
Ishiguro, M. 2008, Icarus, 193, 96
Ishiguro, M., Hanayama, H., Hasegawa, S., et al. 2011, ApJ, 741, L24
Ivezic, Z. et al. 2001, AJ, 122, 5, 2749
Jewitt, D. et al., Weaver, H., Agarwal, J., Mutchler, M., & Drahus, M. 2010, Nature, 467, 817
Jewitt, D. 2011, arXiv:1112.5220
Joye, W. A. & Mandel, E. 2003, Astronomical Data Analysis Software and Systems XII ASP
Conference Series, 295, 489
Landolt, A. U. 1992, AJ, 104, 1, 340
Licandro, J., Campins, H., Kelley, M., et al. 2011, A&A, 525, A34
-- 9 --
Mink, D. 1997, Astronomical Data Analysis Software and Systems VI, A.S.P. Conference series,
125, 249
Moreno, F. et al. 2010, ApJ, 718, L132
Nesvorn´y, D., Morbidelli, A., Vokrouhlick´y, D., Bottke, W. F. & Broz, M. 2001, Icarus, 157, 155
Noguchi, T., Nakamura, T., Kimura, M., et al. 2011, Meteoritics and Planetary Science Supplement,
74, 5117
Sasaki, S., Nakamura, K., Hamabe, Y., Kurahashi, E., & Hiroi, T. 2001, Nature, 410, 6828, 555
Smith, J. A. et al. 2008, AJ, 123, 4, 2121
Snodgrass, C. et al. 2010, Nature, 467, 814
Zappal`a, V., Bendjoya, Ph., Cellino, A., Farinella, & P., Froeschl´e. 1995, Icarus, 116, 291
This preprint was prepared with the AAS LATEX macros v5.2.
-- 10 --
Fig. 1. -- Masking of bright sources on the image. (a) Original image. (b) Resultant image after
removing background stars. Circular masks with zero-values whose sizes were taken in accordance
with brightness of each stars were put to the location of stars.
-- 11 --
Fig. 2. -- Images of dust tail. These are composite images of masked frames in three different
waveband. P/2010 A2 at (a) g′-band, (b) Rc-band, (c) Ic-band
.
-- 12 --
(a)
Cut profiles of dust tail
g'-band
Rc-band
Ic band
20
15
10
5
0
)
U
D
A
(
y
t
i
s
n
e
t
n
I
-20
-15
-10
0
-5
5
Angular distance(")
(b)
10
15
20
Fig. 3. -- (a) Region around the dust tail of P/2010 A2 where the flux profile is examined. (b) Cut
profile of the dust tail showing averaged flux distribution perpendicular to the line of the dust tail.
Relative reflectance can be obtained by comparing total fluxes and solar indices.
-- 13 --
Flora : S-type
P/2010 A2
1.8
1.6
1.4
1.2
1
0.8
400
500
600
700
800
900
Wavelength(nm)
(a)
Flora : Sq-type
P/2010 A2
1.8
1.6
1.4
1.2
1
e
c
n
a
t
c
e
l
f
e
R
e
c
n
a
t
c
e
l
f
e
R
Flora : C-type
P/2010 A2
1.8
1.6
1.4
1.2
1
0.8
400
500
600
700
800
900
Wavelength(nm)
(b)
H5 Chondrites
H6 Chondrites
L4 Chondrites
L5 Chondrites
L6 Chondrites
LL3 Chondrites
LL6 Chondrites
P/2010 A2
1.8
1.6
1.4
1.2
1
e
c
n
a
t
c
e
l
f
e
R
e
c
n
a
t
c
e
l
f
e
R
0.8
400
500
600
700
800
900
0.8
400
500
600
700
800
900
Wavelength(nm)
(c)
Wavelength(nm)
(d)
Fig. 4. -- Reflectances of P/2010 A2 and asteroids representing typical taxonomic types of Flora
family members. Reflectances are normalized at g′-band (470nm) here. (a) S-type asteroids. (b)
C-type asteroids. (c) Sq-type asteroids. For each taxonomic type, average spectra with standard
deviation are demonstrated and they are observation results of SMASS II, Phase II Small Main-Belt
Asteroid Spectroscopic Survey (Bus & Binzel 2002a). (d) Ordinary chondrites. Note that error bar
was indicated for reflectance of H5 chondrites, which shows the most similar spectral property to
that of P/2010 A2.
-- 14 --
Table 1: Relative reflectances of P/2010 A2
Filter band Center wavelength(nm) Relative reflectancea Errorb
-
0.046
0.061
1.000
1.096
1.131
470
650
800
g′
Rc
Ic
aReflectances are normalized at g′-band.
bError of g′-band reflectance was regarded as zero, because we calculated relative reflectances directly from color
indices, (g′
− Rc) and (g′
− Ic).
|
1511.00543 | 2 | 1511 | 2015-11-09T11:18:22 | PAMELA's measurements of geomagnetically trapped and albedo protons | [
"astro-ph.EP",
"physics.space-ph"
] | Data from the PAMELA satellite experiment were used to perform a detailed measurement of under-cutoff protons at low Earth orbits. On the basis of a trajectory tracing approach using a realistic description of the magnetosphere, protons were classified into geomagnetically trapped and re-entrant albedo. The former include stably-trapped protons in the South Atlantic Anomaly, which were analyzed in the framework of the adiabatic theory, investigating energy spectra, spatial and angular distributions; results were compared with the predictions of the AP8 and the PSB97 empirical trapped models. The albedo protons were classified into quasi-trapped, concentrating in the magnetic equatorial region, and un-trapped, spreading over all latitudes and including both short-lived (precipitating) and long-lived (pseudo-trapped) components. Features of the penumbra region around the geomagnetic cutoff were investigated as well. PAMELA observations significantly improve the characterization of the high energy proton populations in near Earth orbits. | astro-ph.EP | astro-ph |
PAMELA's measurements of geomagnetically trapped and albedo
protons
A. Bruno1,∗, O. Adriani2,3, G. C. Barbarino4,5, G. A. Bazilevskaya6, R. Bellotti1,7,
M. Boezio8, E. A. Bogomolov9, M. Bongi2,3, V. Bonvicini8, S. Bottai3, U. Bravar10,
F. Cafagna7, D. Campana5, R. Carbone8, P. Carlson11, M. Casolino12,13, G. Castellini14,
E. C. Christian15, C. De Donato12,17, G. A. de Nolfo15, C. De Santis12,17, N. De Simone12,
V. Di Felice12,18, V. Formato8,19, A. M. Galper16, A. V. Karelin16, S. V. Koldashov16,
S. Koldobskiy16, S. Y. Krutkov9, A. N. Kvashnin6, M. Lee10, A. Leonov16, V. Malakhov16,
L. Marcelli12,17, M. Martucci17,20, A. G. Mayorov16, W. Menn21, M. Merg`e12,17,
V. V. Mikhailov16, E. Mocchiutti8, A. Monaco1,7, N. Mori2,3, R. Munini8,19, G. Osteria5,
F. Palma12,17, B. Panico5, P. Papini3, M. Pearce11, P. Picozza12,17, M. Ricci20,
S. B. Ricciarini3,14, J. M. Ryan10, R. Sarkar22,23, V. Scotti4,5, M. Simon21, R. Sparvoli12,17,
P. Spillantini2,3, S. Stochaj24, Y. I. Stozhkov6, A. Vacchi8, E. Vannuccini3, G. I. Vasilyev9,
S. A. Voronov16, Y. T. Yurkin16, G. Zampa8, N. Zampa8, and V. G. Zverev16.
1 Department of Physics, University of Bari "Aldo Moro", I-70126 Bari, Italy.
2 Department of Physics and Astronomy, University of Florence, I-50019 Sesto Fiorentino,
Florence, Italy.
3 INFN, Sezione di Florence, I-50019 Sesto Fiorentino, Florence, Italy.
4 Department of Physics, University of Naples "Federico II", I-80126 Naples, Italy.
5 INFN, Sezione di Naples, I-80126 Naples, Italy.
6 Lebedev Physical Institute, RU-119991 Moscow, Russia.
7 INFN, Sezione di Bari, I-70126 Bari, Italy.
8 INFN, Sezione di Trieste, I-34149 Trieste, Italy.
9 Ioffe Physical Technical Institute, RU-194021 St. Petersburg, Russia.
10 Space Science Center, University of New Hampshire, Durham, NH, USA.
11 KTH, Department of Physics, and the Oskar Klein Centre for Cosmoparticle Physics,
AlbaNova University Centre, SE-10691 Stockholm, Sweden.
-- 2 --
12 INFN, Sezione di Rome "Tor Vergata", I-00133 Rome, Italy.
13 RIKEN, Advanced Science Institute, Wako-shi, Saitama, Japan.
14 IFAC, I-50019 Sesto Fiorentino, Florence, Italy.
15 Heliophysics Division, NASA Goddard Space Flight Ctr, Greenbelt, MD, USA.
16 National Research Nuclear University MEPhI, RU-115409 Moscow, Russia.
17 Department of Physics, University of Rome "Tor Vergata", I-00133 Rome, Italy.
18 Agenzia Spaziale Italiana (ASI) Science Data Center, I-00133 Rome, Italy.
19 Department of Physics, University of Trieste, I-34147 Trieste, Italy.
20 INFN, Laboratori Nazionali di Frascati, I-00044 Frascati, Italy.
21 Department of Physics, Universitat Siegen, D-57068 Siegen, Germany.
22 Indian Centre for Space Physics, 43 Chalantika, Kolkata 700084, West Bengal, India.
23 Previously at INFN, Sezione di Trieste, I-34149 Trieste, Italy.
24 Electrical and Computer Engineering, New Mexico State University, Las Cruces, NM,
USA.
Received
;
accepted
*Corresponding author. E-mail address: [email protected].
-- 3 --
ABSTRACT
Data from the PAMELA satellite experiment were used to perform a detailed
measurement of under-cutoff protons at low Earth orbits. On the basis of a
trajectory tracing approach using a realistic description of the magnetosphere,
protons were classified into geomagnetically trapped and re-entrant albedo. The
former include stably-trapped protons in the South Atlantic Anomaly, which were
analyzed in the framework of the adiabatic theory, investigating energy spectra,
spatial and angular distributions; results were compared with the predictions
of the AP8 and the PSB97 empirical trapped models. The albedo protons were
classified into quasi-trapped, concentrating in the magnetic equatorial region, and
un-trapped, spreading over all latitudes and including both short-lived (precipi-
tating) and long-lived (pseudo-trapped ) components. Features of the penumbra
region around the geomagnetic cutoff were investigated as well. PAMELA ob-
servations significantly improve the characterization of the high energy proton
populations in near Earth orbits.
1.
Introduction
The radiation environment in Earth's vicinity constitutes a well-known hazard for the
space missions. Major sources include large solar particle events and the Van Allen belts,
consisting of intense fluxes of energetic charged particles experiencing long-term magnetic
trapping. Specifically, the inner belt is mainly populated by protons, mostly originated
by the decay of albedo neutrons according to the CRAND mechanism (Walt 1994). A
standard description of such an environment is provided by the AP8 empirical model
(Sawyer & Vette 1976), based on data from satellite experiments in the 1960s and early
-- 4 --
1970s. Recently, significant improvements (Meffert & Gussenhoven 1994; Huston & Pfitzer
1998; Heynderickx et al. 1999; Xapsos, et al. 2002) have been made thanks to the data from
new spacecrafts (Gussenhoven et al. 1993, 1995; Looper et al. 1996, 1998; Huston et al.
1996). Nevertheless, the modeling of the trapped environment is still incomplete, with
largest uncertainties affecting the high energy fluxes in the inner zone and the South
Atlantic Anomaly (SAA), where the inner belt makes its closest approach to the Earth1.
In addition, the magnetospheric radiation includes populations of albedo protons,
originated by the collisions of Cosmic-Rays (CRs) from interplanetary space on the
atmosphere (Treiman 1953). A quasi-trapped component concentrates in the equatorial
region and presents features similar to those of radiation belt protons, but with limited
lifetimes and much less intense fluxes (Moritz 1972; Hovestadt et al. 1972; Fiandrini et al.
2004). An un-trapped component spreads over all latitudes (Alcaraz et al. 2000; Bidoli et al.
2003) including the "penumbra" region around the geomagnetic cutoff, where particles of
both cosmic and atmospheric origin are present (Cooke et al. 1991).
New accurate measurements of the CR radiation at low Earth orbits have been
performed by the PAMELA experiment (Adriani et al. 2014). This paper reports the
observations of the geomagnetically trapped and re-entrant albedo protons.
2. Data analysis
PAMELA is a space-based experiment designed for a precise measurement of charged
CRs in the energy range from some tens of MeV up to several hundreds of GeV. The
1The SAA is a consequence of the tilt (∼10 deg) between the magnetic dipole axis of the
Earth and its rotational axis, and of the offset (∼500 km) between the dipole and the Earth
centers.
-- 5 --
Resurs-DK1 satellite, which hosts the apparatus, has a semi-polar (70 deg inclination) and
elliptical (350÷610 km altitude) orbit. The spacecraft is 3-axis stabilized; its orientation
is calculated by an onboard processor with an accuracy better than 1 deg. Particle
directions are measured with a high angular resolution (< 2 deg). Details about apparatus
performance, proton selection, detector efficiencies and experimental uncertainties can be
found elsewhere (see e.g. Adriani et al. (2013)). The data set analyzed in this work includes
protons collected by PAMELA between 2006 July and 2009 September.
2.1. Particle classification
Trajectories of all detected down-going protons were reconstructed in the Earth's
magnetosphere using a tracing program based on numerical integration methods
(Smart & Shea 2000, 2005), and implementing the IGRF11 (Finlay et al. 2010) and
the TS05 (Tsyganenko & Sitnov 2005) as internal and external geomagnetic models.
Trajectories were propagated back and forth from the measurement location, and traced
until: they reached the model magnetosphere boundaries (galactic protons); or they
intersected the absorbing atmosphere limit, which was assumed at an altitude2 of 40 km
(re-entrant albedo protons); or they performed more than 106/R2 steps3, where R is the
particle rigidity in GV, for both propagation directions (geomagnetically trapped protons).
Trapped trajectories were verified to fulfil the adiabatic conditions (Adriani et al. 2015a),
2Such a value approximately corresponds to the mean production altitude for albedo
protons.
3Since the program uses a dynamic variable step length, which is of the order of 1% of
a particle gyro-distance in the magnetic field, such a criterion ensures that at least 4 drift
cycles around the Earth were performed.
-- 6 --
in particular the hierarchy of temporal scales: ωgyro ≫ ωbounce ≫ ωdrif t, where ωgyro, ωbounce
and ωdrif t are the frequencies associated with gyration, bouncing and drift motions.
Albedo protons were classified into quasi-trapped and un-trapped. The former have
trajectories similar to those of stably-trapped, but are originated and re-absorbed by the
atmosphere during a time larger than a bounce period (up to several tens of s). The latter
include both a short-lived component of protons precipitating into the atmosphere within a
bounce period (. 1s), and a long-lived (pseudo-trapped ) component with rigidities near the
geomagnetic cutoff (penumbra region), characterized by a chaotic motion (non-adiabatic
trajectories). Further details, including distributions of lifetimes and production/absorption
points on the atmosphere, can be found in Adriani et al. (2015b).
2.2. Flux calculation
Proton fluxes were derived by assuming an isotropic flux distribution in all the
explored regions except the SAA. In this case, fluxes are significantly anisotropic due to the
interaction with the Earth's atmosphere, and thus the gathering power of the apparatus
(Sullivan 1971) depends on the spacecraft orientation with respect to the geomagnetic
field. Consequently, a PAMELA effective area (cm2) was evaluated as a function of particle
energy E, local pitch angle α and satellite orientation Ψ:
H(E, α, Ψ) =
sinα
2π Z 2π
0
dβ [A(E, θ, φ) · cosθ] ,
(1)
where β is the gyro-phase angle, θ=θ(α, β, Ψ) and φ=φ(α, β, Ψ) are respectively the zenith
and the azimuth angle describing particle direction in the PAMELA frame4, and A(E, θ, φ)
is the apparatus response function. The effective area was evaluated with accurate Monte
4The PAMELA frame has the origin in the center of the spectrometer cavity; the Z axis
is directed along the main axis of the apparatus, toward incoming particles; the Y axis is
-- 7 --
Fig. 1. -- Proton integral fluxes (m−2s−1sr−1) as a function of equatorial pitch angle αeq and
McIlwain's L-shell, for different kinetic energy bins (see the labels). Results for the various
components are reported (from left to right): stably-trapped, quasi-trapped, un-trapped and
the total sample. See the text for details.
Carlo simulations based on integration methods (Sullivan 1971). Finally, in order to account
for effects due to the large particle gyro-radius (up to several hundreds of km), trapped
fluxes were evaluated by shifting measured protons (L, B, Beq) to corresponding guiding
center positions. Further details can be found in Adriani et al. (2015a).
directed opposite to the main direction of the magnetic field inside the spectrometer; the X
axis completes a right-handed system.
-- 8 --
3. Results
Figure 1 shows the fluxes of under-cutoff protons as a function of equatorial pitch angle
αeq and McIlwain's L-shell, integrated over different kinetic energy bins. The first column
reports the results for stably-trapped protons, concentrating in the SAA at PAMELA
altitudes. Constrained by the spacecraft orbit, the covered phase-space region varies with
the magnetic latitude. In particular, PAMELA can observe equatorial mirroring protons
only for L-shell values up to ∼1.18 RE, and measured distributions are strips of limited
width parallel to the "drift loss cone", which delimits the αeq range for which stable
magnetic trapping does not occur. Fluxes exhibit strong angular and radial dependencies.
PAMELA is able to measure trapped spectra up to their highest energies (about 4 GeV)
(Adriani et al. 2015a). For a comparison, Figure 1 also reports the fluxes for quasi- and
un-trapped components. In this case, measured maps5 result from the superposition of
distributions corresponding to regions characterized by a different local (or bounce) loss
cone value. Fluxes are quite isotropic except in the SAA, where distributions are similar to
those of stably-trapped protons (Adriani et al. 2015b).
Figure 2 compares PAMELA geomagnetically trapped results with the predictions
from two empirical models available in the same energy and altitude ranges: the AP8
(Sawyer & Vette 1976) unidirectional (or UP8 (Daly & Evans 1996)) model for solar
minimum conditions, and the SAMPEX/PET PSB97 model (Heynderickx et al. 1999).
Data were derived by using the SPENVIS web-tool (Heynderickx et al. 2002). In general,
the UP8 model significantly overestimates PAMELA observations, while a better agreement
can be observed with the PSB97 model. However, PAMELA fluxes do not show the spectral
structures present in the PSB97 predictions.
5Note that the un-trapped flux suppression at highest energy and L bins is due to the
geomagnetic cut (R < 10/L3) used for selecting adiabatic trajectories.
-- 9 --
Fig. 2. -- PAMELA trapped proton energy spectrum for sample αeq and L-shell values,
compared with the predictions from the UP8-min (Sawyer & Vette 1976; Daly & Evans 1996)
and the PSB97 (Heynderickx et al. 1999) models (from SPENVIS (Heynderickx et al. 2002)).
Albedo fluxes were mapped using the Altitude Adjusted Corrected Geomagnetic
(AACGM) coordinates (Heres & Bonito 2007), developed to provide a more realistic
description of high latitude regions, by accounting for the multipolar geomagnetic field.
Figure 3 shows the spectra of the various albedo components outside the SAA (B>0.23 G)
measured at different latitudes, along with the galactic component. Fluxes were averaged
over longitudes. Quasi-trapped protons are limited to low latitudes and to energies below
∼ 8 GeV; their fluxes smoothly decrease with increasing latitude and energy. Conversely,
the precipitating component spreads to higher latitudes, with spectra extending up to ∼10
GeV. Finally, pseudo-trapped protons concentrate at highest latitudes and energies (up to
∼ 20 GeV), with a peak in the penumbra originated by large gyro-radius (102 ÷ 103 km)
effects.
Features of the penumbra region are investigated in Figure 4, where the fraction
of galactic over total (galactic + albedo) protons is displayed as a function of particle
rigidity and AACGM latitude (left panels); for a comparison, distributions as a function
Λ0 <
< 5
1
-
]
2
m
r
s
s
V
e
G
[
0.1
10
Kin. Energy [GeV]
1
Λ
15 <
< 20
x
u
F
l
1
-
]
2
m
r
s
-- 10 --
Λ5 <
< 10
1
-
]
2
m
r
s
s
V
e
G
[
0.1
10
Kin. Energy [GeV]
1
Λ
20 <
< 25
x
u
F
l
1
-
]
2
m
r
s
0.1
10
Kin. Energy [GeV]
1
Λ
30 <
< 35
0.1
10
Kin. Energy [GeV]
1
Λ
45 <
< 50
0.1
10
Kin. Energy [GeV]
1
Λ
60 <
< 65
0.1
10
Kin. Energy [GeV]
1
0.1
10
Kin. Energy [GeV]
1
Λ
35 <
< 40
0.1
10
Kin. Energy [GeV]
1
Λ
50 <
< 55
0.1
10
Kin. Energy [GeV]
1
Λ
65 <
< 70
0.1
10
Kin. Energy [GeV]
1
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
410
310
210
10
1
-110
-210
-310
410
310
210
10
1
-110
-210
-310
s
V
e
G
[
x
u
F
l
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
410
310
210
10
1
-110
-210
-310
410
310
210
10
1
-110
-210
-310
s
V
e
G
[
x
u
F
l
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
1
-
]
2
m
r
s
s
V
e
G
[
x
u
F
l
410
310
210
10
1
-110
-210
-310
Λ
10 <
< 15
0.1
10
Kin. Energy [GeV]
1
Λ
25 <
< 30
0.1
10
Kin. Energy [GeV]
1
Λ
40 <
< 45
0.1
10
Kin. Energy [GeV]
1
Λ
55 <
< 60
0.1
10
Kin. Energy [GeV]
1
Fig. 3. -- Differential energy spectra (GeV−1 m−2 s−1 sr−1) outside the SAA for different
AACGM latitude Λ bins. Results for the several proton populations are shown: quasi-
trapped (violet), precipitating (green), pseudo-trapped (red) and galactic (blue). Lines are
to guide the eye.
of McIlwain's L-shell are also shown (right panels). The penumbra was identified as the
region where both albedo and galactic proton trajectories were reconstructed. The black
curves denote a fit of points with an equal percentage of the two components, while the
-- 11 --
10
]
V
G
[
y
t
i
d
g
R
i
i
1
0
10
]
%
[
n
o
i
t
c
a
r
f
c
i
t
c
a
l
a
G
100
90
80
70
60
50
40
30
20
10
0
×
×
11.62
11.62
cos
cos
Λ(4
Λ(4
) - 0.215 GV
) - 0.215 GV
30
20
50
AACGM latitude [deg]
40
1
Rigidity [GV]
10
100
90
80
70
60
50
40
30
20
10
G
a
l
a
c
t
i
c
f
r
a
c
t
i
o
n
[
%
]
60
70
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
Λ
= 1.5
= 1.5
= 4.5
= 4.5
= 7.5
= 7.5
= 10.5
= 10.5
= 13.5
= 13.5
= 16.5
= 16.5
= 19.5
= 19.5
= 22.5
= 22.5
= 25.5
= 25.5
= 28.5
= 28.5
= 31.5
= 31.5
= 34.5
= 34.5
= 37.5
= 37.5
= 40.5
= 40.5
= 43.5
= 43.5
= 46.5
= 46.5
= 49.5
= 49.5
= 52.5
= 52.5
= 55.5
= 55.5
= 58.5
= 58.5
= 61.5
= 61.5
= 64.5
= 64.5
10
]
V
G
[
y
t
i
d
g
R
i
i
1
1
]
%
[
n
o
i
t
c
a
r
f
c
i
t
c
a
l
a
G
100
90
80
70
60
50
40
30
20
10
0
14.3 / L
14.3 / L
GV2
GV2
2
2
14.14 / L
- 0.187 GV
14.14 / L
- 0.187 GV
2
3
L-shell [R
]
E
4
5
6
1
Rigidity [GV]
10
100
90
80
70
60
50
40
30
20
10
G
a
l
a
c
t
i
c
f
r
a
c
t
i
o
n
[
%
]
L = 0.99
L = 0.99
L = 1.07
L = 1.07
L = 1.16
L = 1.16
L = 1.25
L = 1.25
L = 1.35
L = 1.35
L = 1.46
L = 1.46
L = 1.58
L = 1.58
L = 1.71
L = 1.71
L = 1.85
L = 1.85
L = 2.00
L = 2.00
L = 2.16
L = 2.16
L = 2.34
L = 2.34
L = 2.53
L = 2.53
L = 2.73
L = 2.73
L = 2.95
L = 2.95
L = 3.19
L = 3.19
L = 3.45
L = 3.45
L = 3.73
L = 3.73
L = 4.03
L = 4.03
L = 4.36
L = 4.36
L = 4.72
L = 4.72
L = 5.10
L = 5.10
L = 5.51
L = 5.51
L = 5.96
L = 5.96
L = 6.44
L = 6.44
Fig. 4. -- Fraction of galactic protons in the penumbra region, as a function of particle
rigidity and AACGM latitude Λ (left) and McIlwain's L-shell (right). See the text for
details.
red line refers to the Stormer vertical cutoff for the PAMELA epoch. Bottom panels report
corresponding rigidity profiles.
4. Summary and Conclusions
PAMELA measurements of energetic (>70 MeV) under-cutoff proton fluxes at low
Earth orbits (350÷610 km) have been presented. The detected proton sample was classified
into geomagnetically trapped and re-entrant albedo on the basis of accurate particle tracing
techniques.
Stably-trapped protons, confined in the SAA at PAMELA altitudes, were investigated
in the framework of the adiabatic theory. PAMELA data extend the observational range
for the trapped radiation down to lower L-shells (∼ 1.1 RE) and up to highest kinetic
-- 12 --
energies (. 4 GeV), significantly improving the description of the low altitude radiation
environment, where current models suffer from the largest uncertainties.
Albedo protons were classified into quasi-trapped and un-trapped : the former consist
of relatively long-lived protons populating the equatorial region, with trajectories similar
to those of stably-trapped; the latter include a short-lived (precipitating) component
spreading over all explored latitudes, along with a long-lived (pseudo-trapped ) component
concentrating near the geomagnetic cutoff and characterized by a chaotic motion
(non-adiabatic trajectories).
PAMELA results significantly enhance the characterization of high energy proton
populations in a wide geomagnetic region, enabling a more precise and complete view of
atmospheric and magnetospheric effects on the CR transport near the Earth.
-- 13 --
REFERENCES
O. Adriani, et al., 2013, ApJ 765:91.05205.
O. Adriani, et al., 2014, Physics Reports, Vol. 544, 4, pp. 323 -- 370,
doi:10.1016/j.physrep.2014.06.003
O. Adriani, et al., 2015a, ApJ 799 L4, doi:10.1088/2041-8205/799/1/L4.
O. Adriani, et al., 2015b, J. Geophys. Res. Space Physics, 120, doi:10.1002/2015JA021019.
J. Alcaraz, et al., 2000, Phys. Lett. B 472:215 -- 226, doi:10.1016/S0370-2693(99)01427-6.
V. Bidoli, et al., 2003, J. Geophys. Res., 108, A5, 1211, doi:l0.1029/2002JA009684.
D. J. Cooke, Humble, J. E., Smart, M. A., et al., 1991, Il Nuovo Cimento, 14C, 213.
E. J. Daly & H. D. R. Evans, 1996, Rad. Meas., vol. 26, no. 3, pp. 363-368.
E. Fiandrini, et al., 2004, J. Geophys. Res., 109, 10214, doi:10.1029/2004JA010394.
C. C. Finlay, et al., 2010, Geophysical Journal International, 183: 1216 -- 1230.
M. S. Gussenhoven, et al., 1993, IEEE Trans. Nucl. Sci., 40, 1450.
M. S. Gussenhoven, et al., 1995, IEEE Trans. Nucl. Sci., 42, 2035.
W. Heres & N. A. Bonito, 2007, Scientific Report AFRL -- RV -- HA -- TR -- 2007 -- 1190.
D. Heynderickx, et al., 2002, SAE Technical Paper 2000-01-2415.
D. Heynderickx, et al., 1999, IEEE Trans. Nucl. Sci., 46, 1475.
D. Hovestadt, et al., 1972, Phys. Rev. Lett. 28, 1340.
S. L. Huston, G. A. Kuck and K. A. Pfitzer, 1996, Geophys. Monogr. Ser., Vol. 97, pp.
119 -- 122.
-- 14 --
S. L. Huston & K. A. Pfitzer, 1998, NASA Contract. Rep. NASA/CR-1998-208593.
M. D. Looper, et al., 1996, Radiat. Meas., 26(6):967-78.
M. D. Looper, J. B. Blake and R. A. Mewaldt, 1998, Adv. Space Res., 21(12):1679-82.
J. D. Meffert & M. S. Gussenhoven, 1994, PL-TR-94-2218, Environ. Res. Pap. 1158, Air
Force Res. Lab., Wright-Patterson Air Force Base, Ohio.
J. Moritz, 1972, Z. Geophys. 38, 701.
D. M. Sawyer & J. I. Vette, 1976, NSSDC/WDC-A-R&S 76-06.
D. F. Smart & M. A. Shea, 2000, Final Report, Grant NAG5 -- 8009, Center for Space
Plasmas and Aeronomic Research, The University of Alabama in Huntsville.
D. F. Smart & M. A. Shea, 2005, Adv. Space Res., 36, 2012 -- 2020.
J. D. Sullivan, 1971, Nucl. Instr. and Meth. 95, 5.
S. B. Treiman, 1953, The Cosmic-Ray Albedo, Phys. Rev. 53, 957,
doi:http://dx.doi.org/10.1103/PhysRev.91.957.
N. A. Tsyganenko & M. I. Sitnov, 2005, J. Geophys. Res., Vol. 110, A03208.
M. A. Xapsos, et al., 2002, IEEE Trans. Nucl. Sci., 49, 2776.
M. Walt, 1994, Introduction to Geomagnetically Trapped Radiation, Cambridge Univ.
Press. (New York).
This manuscript was prepared with the AAS LATEX macros v5.2.
|
1909.05851 | 3 | 1909 | 2019-10-14T17:59:08 | Initial characterization of interstellar comet 2I/Borisov | [
"astro-ph.EP",
"astro-ph.SR"
] | Interstellar comets penetrating through the Solar System had been anticipated for decades. The discovery of asteroidal-looking 'Oumuamua was thus a huge surprise and a puzzle. Furthermore, the physical properties of the 'first scout' turned out to be impossible to reconcile with Solar System objects, challenging our view of interstellar minor bodies. Here, we report the identification and early characterization of a new interstellar object, which has an evidently cometary appearance. The body was discovered by Gennady Borisov on 30 August 2019 UT and subsequently identified as hyperbolic by our data mining code in publicly available astrometric data. The initial orbital solution implies a very high hyperbolic excess speed of ~32 km/s, consistent with 'Oumuamua and theoretical predictions. Images taken on 10 and 13 September 2019 UT with the William Herschel Telescope and Gemini North Telescope show an extended coma and a faint, broad tail. We measure a slightly reddish colour with a g'-r' colour index of 0.66 +/- 0.01 mag, compatible with Solar System comets. The observed morphology is also unremarkable and best explained by dust with a power-law size-distribution index of -3.7 +/- 1.8 and a low ejection speed (44 +/- 14 m/s for $\beta$ = 1 particles, where $\beta$ is the ratio of the solar gravitational attraction to the solar radiation pressure). The nucleus is probably ~1 km in radius, again a common value among Solar System comets, and has a negligible chance of experiencing rotational disruption. Based on these early characteristics, and putting its hyperbolic orbit aside, 2I/Borisov appears indistinguishable from the native Solar System comets. | astro-ph.EP | astro-ph | Initial characterization of interstellar comet 2I/Borisov
Piotr Guzik1*, Michał Drahus1*, Krzysztof Rusek2, Wacław Waniak1, Giacomo Cannizzaro3,4,
Inés Pastor-Marazuela5,6
1 Astronomical Observatory, Jagiellonian University, Kraków, Poland
2 AGH University of Science and Technology, Kraków, Poland
3 SRON, Netherlands Institute for Space Research, Utrecht, the Netherlands
4 Department of Astrophysics/IMAPP, Radboud University, Nijmegen, the Netherlands
5 Anton Pannekoek Institute for Astronomy, University of Amsterdam, Amsterdam, the Netherlands
6 ASTRON, Netherlands Institute for Radio Astronomy, Dwingeloo, the Netherlands
* These authors contributed equally to this work; email: [email protected], [email protected]
Interstellar comets penetrating through the Solar System had been anticipated for
decades1,2. The discovery of asteroidal-looking 'Oumuamua3,4 was thus a huge surprise
and a puzzle. Furthermore, the physical properties of the 'first scout' turned out to be
impossible to reconcile with Solar System objects4 -- 6, challenging our view of interstellar
minor bodies7,8. Here, we report the identification and early characterization of a new
interstellar object, which has an evidently cometary appearance. The body was
discovered by Gennady Borisov on 30 August 2019 UT and subsequently identified as
hyperbolic by our data mining code in publicly available astrometric data. The initial
orbital solution implies a very high hyperbolic excess speed of ~32 km s−1, consistent with
'Oumuamua9 and theoretical predictions2,7. Images taken on 10 and 13 September 2019
UT with the William Herschel Telescope and Gemini North Telescope show an extended
coma and a faint, broad tail. We measure a slightly reddish colour with a g′ -- r′ colour
index of 0.66 ± 0.01 mag, compatible with Solar System comets. The observed
morphology is also unremarkable and best explained by dust with a power-law size-
distribution index of -- 3.7 ± 1.8 and a low ejection speed (44 ± 14 m s−1 for β = 1 particles,
where β is the ratio of the solar gravitational attraction to the solar radiation pressure).
The nucleus is probably ~1 km in radius, again a common value among Solar System
comets, and has a negligible chance of experiencing rotational disruption. Based on these
early characteristics, and putting its hyperbolic orbit aside, 2I/Borisov appears
indistinguishable from the native Solar System comets.
On 8 September 2019 at 04:15 UT, we were alerted by our software Interstellar Crusher (see
Methods) of a possible new hyperbolic object gb00234. Within less than 4 d, the orbit became
reliable enough to trigger the first announcements10,11, and subsequently, the body received an
official name 2I/Borisov. As of 20 September 2019 at 12:00 UT, 447 published astrometric
positions collected over a 21.1 d interval11 -- 15 are demonstrably incompatible with a parabolic
orbit (Fig. 1). In the absence of non-gravitational forces, the residuals show a very strong
systematic trend and reach up to 20 arcsec. Inclusion of non-gravitational forces greatly
reduces the residuals, but a small systematic trend is still present. More importantly, the
resulting non-gravitational accelerations A1 = 7.43 ± 0.07 × 10−4 au d−2 and A2 = −1.83 ±
0.01 × 10−4 au d−2 given at 1 au from the Sun (see Methods) are implausibly high, exceeding
the largest measured non-gravitational accelerations of comets16 by two to three orders of
magnitude and comparable to the sunward gravitational acceleration at 1 au. However, the
data are accurately fitted by an unrestricted, purely gravitational solution (Fig. 1) that implies
a strongly hyperbolic orbit with an eccentricity of 3.38 ± 0.02 (Table 1). This strong
hyperbolicity cannot be attributed to gravitational perturbations from the Solar System's
planets because the body travels from a direction far from the ecliptic plane. Thus, the only
viable explanation is the arrival from outside the Solar System. The huge eccentricity together
with a moderate perihelion distance of 2.012 ± 0.004 au (Table 1) imply a hyperbolic excess
speed of ~32 km s−1. The body entered the Solar System from a direction ~75° away from the
Solar apex with the asymptotic radiant at J2000.0 right ascension (RA) = 02 h 12 m and
declination (dec) = 59.4° in the constellation of Cassiopeia. For the most up-to-date orbital
parameters, readers are referred to the online databases of the Minor Planet Center or Jet
Propulsion Laboratory.
Table 1 Hyperbolic heliocentric orbital elements of 2I/Borisov calculated for the
osculation epoch 2019 September 20.0 TT
T
e
q
ω
Ω
i
2019 Dec. 8.42 ± 0.11 TT
3.3790 ± 0.020
2.0119 ± 0.0044 au
209.001 ± 0.100° (J2000.0)
308.195 ± 0.040° (J2000.0)
44.004 ± 0.041° (J2000.0)
Fig. 1 Astrometric residuals of 2I/Borisov calculated for three different orbital
solutions. a,b, The residuals calculated for a parabolic solution without non-gravitational
forces. c,d, The residuals for a parabolic solution with non-gravitational forces. e,f, The
residuals for an unrestricted solution without non-gravitational forces. Residuals are presented
separately in RA (a,c,e) and dec (b,d,f). Black symbols denote data used in the computation
and red symbols denote rejected outliers. r.m.s., root mean squared.
We observed this object using the 4.2 m William Herschel Telescope (WHT) on La Palma
and the 8.2 m Gemini North Telescope at Maunakea in the Sloan Digital Sky Survey (SDSS)
g′ and r′ bands17. WHT data were obtained with the Auxiliary-port CAMera (ACAM) on 10
September 2019 at 05:38 UT and on 13 September 2019 at 05:47 UT (observation mid-
points). The first set comprises ten sidereal-tracked 60 s exposures, of which five were
obtained in the g′ band and five in the r′ band, whereas the second set contains 40 sidereal-
tracked 20 s exposures, 20 obtained in g′ and 20 in r′. Gemini data were collected with the
Gemini Multi-Object Spectrograph (GMOS-N) on 10 September 2019 at 14:57 UT (mid-
point) with non-sidereal tracking and comprise four g′-band and four r′-band exposures taken
with 60 s integration time. The datasets were obtained at low elevation (22° to 31°) in
morning twilight (solar elevation from −19° to −12°). At the time of the observations, the
helio- and geocentric distances of 2I/Borisov were equal to 2.8 and 3.4 au, respectively, the
phase angle was ~15° and the apparent motion was ~75 arcsec h−1. The images were corrected
for overscan, bias and flatfield in the standard fashion, and then a global background level
was subtracted from each frame.
In Fig. 2 we show median-stacked g′-band and r′-band images from Gemini, which have a
better signal-to-noise ratio than the WHT images. The latter are presented in Supplementary
Fig. 1. The images reveal an extended coma and a broad, short tail emanating in roughly
antisolar direction. We see no clear difference in morphology in the two bands. The comet
was measured photometrically in each individual exposure from the two telescopes and two
nights. We consistently used a 5,000 km (~2 arcsec) radius photometric aperture and
determined the brightness against background stars available in the SDSS photometric
catalogue18. As a result, we obtained dataset-averaged AB magnitudes g′ = 19.38 ± 0.01 and
r′ = 18.71 ± 0.01 for the WHT observations on 10 September 2019 at 05:38 UT,
g′ = 19.38 ± 0.01 and r′ = 18.72 ± 0.01 for the Gemini observations on 10 September 2019 at
14:57 UT, and g′ = 19.32 ± 0.02 and r′ = 18.67 ± 0.01 for the WHT observations on 13
September 2019 at 05:47 UT. Photometric magnitudes can be used to estimate the nucleus
size. Following two independent approaches, we have found that the nucleus of 2I/Borisov is
most likely ~1 km in radius (see Methods). However, it should be noted that these approaches
are inherently very uncertain. Our photometric measurements give a consistent g′ -- r′ colour
index, with an average value of 0.66 ± 0.01 mag. The colour is slightly redder than the solar
g′ -- r′ = 0.45 ± 0.02 mag19 and implies a positive spectral slope S′ ~ 12.5% per 100 nm (see
Methods), in good agreement with an independent spectroscopic determination20. Within the
errors, the same colour is obtained for other photometric apertures as well (we made
measurements for the apertures ranging from 3,000 to 15,000 km in radius), implying a
uniform colour of the coma. This, and the similarity of morphology in both bands, are both
indicative of dust-dominated activity in our data. Monte Carlo modelling of the object's dust
environment with our established code21 has revealed a power-law particle size-distribution
exponent of -- 3.7 ± 1.8 and a 44 ± 14 m s−1 ejection speed applicable to β = 1 particles (where β
is the ratio of the solar gravitational attraction to the solar radiation pressure; see Methods).
Fig. 2 Median-stacked images of 2I/Borisov from Gemini North. The left panel shows the
image in the g′ band and the right panel shows the image in the r′ band. Both panels subtend
1.0 × 1.0 arcmin and were scaled logarithmically. Arrows show the directions of north (N) and
east (E), the projected antisolar vector ( -- ⊙) and the negative of the orbital velocity vector
( -- V). The pixel scale is 0.16 arcsec px−1 and the seeing was 1.8 arcsec.
The dynamical properties and morphology of 2I/Borisov make it clear that the body is the first
certain case of an interstellar comet, and the second known interstellar minor body identified
in the Solar System (after 'Oumuamua3,4). Evidently, the extended coma and the broad tail
stand in stark contrast with the purely asteroidal appearance of 'Oumuamua. The estimated
nucleus size is common among Solar System's comets22,23. Adopting the formalism of
rotational disruption probability24 and assuming typical properties of Solar System comets, we
have estimated the probability of rotational disruption during the Solar System flyby to be
smaller than 1% (see Methods). The measured colour is consistent with the colours of the
Solar System's comets25 -- 27 and falls only slightly redwards of the median and average values
of the observed g′ -- r′ distribution (Fig. 3). The same similarity can be noticed for the dust
coma parameters28,29. These facts are remarkable in and of themselves, and especially
remarkable after 'Oumuamua, the multiple peculiarities of which4 -- 6 prompted us to rethink
our entire view of the nature of interstellar interlopers7. However, 2I/Borisov appears
completely similar to the native Solar System's comets.
Fig. 3 Colour of 2I/Borisov in the context of Solar System comets. The histogram shows
the g′ -- r′ colour index distribution for 60 comets25 -- 27 (long-period, Jupiter-family and active
Centaurs). It was calculated from the colour indices reported in the Johnson -- Cousins
(UBVRCIC) and the original SDSS (ugriz) photometric systems using standard transformation
formulae36,45. Whenever available, multiple measurements of a single object were filtered and
combined to minimize the colour uncertainty. The sample has an r.m.s. error of g′ -- r′ equal to
0.036 mag and the error does not exceed 0.075 mag for any object (the latter criterion resulted
in the rejection of three comets). Dashed lines show the median (blue) and average (green)
values of the distribution, and the red arrow indicates the measured g′ -- r′ colour index of
2I/Borisov.
Comet 2I/Borisov was discovered on its way to perihelion (8 December 2019 UT at 2.0 au)
and before the closest approach to Earth (28 December 2019 UT at 1.9 au); thus, the overall
visibility will be gradually improving. The body is destined for an intensive observing
campaign lasting many months (Supplementary Fig. 2) that will allow us to gain
groundbreaking insights into the physical properties of interstellar comets and exosolar
planetary systems in general. This discovery is in line with the detection statistic of one
interstellar object per year proposed by several authors30,31 after the discovery of 'Oumuamua,
which is an order of magnitude higher than the most optimistic pre-'Oumuamua estimates32. It
also shows that cometary nature of these bodies might be a common characteristic, in
agreement with the original expectations1,2. More discoveries are expected in the near future
thanks to the Large Synoptic Survey Telescope.
Methods
Interstellar Crusher
Interstellar Crusher is a custom Python3 code running on Windows Subsystem for Linux. It
continuously monitors the Possible Comet Confirmation Page and computes orbits of newly
discovered minor bodies using Bill Gray's Find_Orb (https://github.com/Bill-
Gray/find_orb/commit/abe3f5847aad4c39f1d82239bf343c876b3d1bd0). Detection of a
possible interstellar object triggers an alarm that is sent via e-mail.
Orbit
We computed the orbit using Bill Gray's Find_Orb with planetary perturbations. The initial
input dataset comprised 447 astrometric positions obtained at 45 different stations that were
publicly available as of 20 September 2019 at 12:00 UT. We computed parabolic and
unrestricted solutions without non-gravitational forces and a parabolic solution with non-
gravitational forces defined according to the standard Marsden -- Sekanina model33. Given the
absence of reliable uncertainties of individual observations, we assumed equal weights.
Astrometric positions with residuals greater than 1.5 arcsec with respect to the unrestricted
solution without non-gravitational forces were iteratively rejected, resulting in a restricted
dataset of 362 measurements. This dataset was used for the final computations, the results of
which are presented in Fig. 1 and Table 1. The reported uncertainties of the orbital elements
were estimated by the least squares method from the astrometric residuals.
Photometry and colour
Individual calibrated exposures were first scrutinized for background stars, cosmic-ray hits
and other artefacts in the comet's photometric aperture. As a result of this procedure, we
rejected one g′-band frame from the Gemini dataset, one g′-band and one r′-band frame from
the first WHT dataset, and accepted all frames from the second WHT dataset. For each
dataset, we identified a set of field stars available in the SDSS photometric catalogue18 that
were brighter than the comet and had a g′ -- r′ colour index between 0.3 and 1.0 mag. We
identified 7 and 9 suitable stars for the first and second WHT dataset, respectively, and 4 stars
for the Gemini dataset. Photometric measurements of the comet were done using a circular
aperture with a 5,000 km (~2.0 arcsec) radius. The stars were measured in a larger, 4.0 arcsec
radius aperture that contained >99% of the flux (determined from the curve of growth). Sky
background level was estimated (as a mean value) in an annular aperture with a radius of 7.0 --
9.0 arcsec for the comet and 5.0 -- 8.0 arcsec for the stars. We masked the regions of the
background aperture contaminated by the comet's tail, faint field stars and cosmic-ray hits.
Centroids were measured from a 1.0 arcsec radius aperture. Differential magnitudes of the
comet were calculated for each individual image and then dataset averaged (with equal
weights) and compared with the SDSS catalogue magnitudes of the reference stars. The
photometric uncertainties were propagated from the individual differential measurements and
SDSS uncertainties of the reference stars.
Colour index (m1 -- m2) can be easily converted to the normalized reflectance slope S′. By
definition
(𝑚(cid:2869) − 𝑚(cid:2870)) − (𝑚(cid:2869) − 𝑚(cid:2870))⊙ = −2.5 log (𝑆(cid:2869)/𝑆(cid:2870))
where (𝑚(cid:2869) − 𝑚(cid:2870))⊙ is the solar colour index, 𝑆(cid:2971) = 1 + 𝑆(cid:4593)(𝜆 − 𝜆(cid:2868))/∆𝜆(cid:2868) is the normalized
reflectance at wavelength 𝜆, ∆𝜆(cid:2868) defines the standard wavelength interval of S′ and 𝜆(cid:2868) is the
normalization wavelength. Using the customary values 𝜆(cid:2868) = 550 nm and ∆𝜆(cid:2868) = 100 nm and
substituting the measured g′ -- r′ colour index of comet 2I/Borisov equal to 0.66 mag along with
the effective filter wavelengths 𝜆(cid:2917)(cid:4593) = 475 nm and 𝜆(cid:2928)(cid:4593) = 630 nm and the solar
g′ -- r′ = 0.45 mag19, we calculated the corresponding slope S′ = 12.5% per 100 nm.
Nucleus size estimation
To estimate the size of the nucleus, we followed a simple approach2 that connects theoretical
sublimation rate of water from a unit surface area34 with an empirical correlation of the
observed total visual magnitudes with the total water production rates35. By solving the
standard energy budget equation34 with an assumed Bond albedo of 4% and emissivity of
100%, we find that the average water sublimation flux from a spherical non-rotating nucleus
is 1.86 × 1026 molecules s−1 km−2 at the heliocentric distance applicable to our observations.
From our Gemini data, we estimate the total visual magnitude to be ~17.5, which is the
asymptotic magnitude from the curve of growth transformed from the SDSS g′ and r′ bands to
the Johnson V band36. Taking into account the geo- and heliocentric distances at the time of
the observation, this magnitude corresponds to the water production rate of
~1027 molecules s−1 according to an empirical relation35. By comparing the theoretical and
observed water production rates, we estimate the area of the sublimating surface to be ~5 km2,
which corresponds to ~1 km radius nucleus with 30% active fraction, assuming negligible
contribution of icy grains in the coma.
The same result is obtained from a comparison of 2I/Borisov with the well-studied comet
Hale -- Bopp in terms of the Afρ parameter37. From our r′-band magnitudes measured in the
5,000 km radius aperture, we calculate Afρ ~ 100 cm, and similar for other aperture sizes. The
Afρ of comet Hale -- Bopp was ~10,000 cm at the heliocentric distance and phase angle
compatible with our observations38. Simple scaling of the ~30 km in radius nucleus of the
latter comet39 suggests the nucleus radius of 2I/Borisov of ~1 km.
Probability of rotational disruption
Rotational disruption occurs when the nucleus rotation rate becomes too high for the self-
gravity and tensile strength to keep the body intact40. The probability of disruption P has been
formulated24 as the ratio of the expected change in the rotation frequency Δω to the total
extent of tolerable frequency regime, limited by the negative and positive critical frequency
ωcrit. Thus
𝑃 =
Δ𝜔
2𝜔(cid:2913)(cid:2928)(cid:2919)(cid:2930)
The two components of this equation can be calculated in a relatively straightforward manner
using standard formulae5,41 -- 44. Substituting in these formulas the estimated nucleus radius of
~1 km and adopting typical properties of Solar System comets24, we find the probability of
rotational disruption of 2I/Borisov to be <1%.
Dust modelling
We modelled the dust environment of 2I/Borisov using a Monte Carlo approach21. The
simulations were done with 2 × 106 power-law distributed dust particles spanning a size range
a from 0.5 μm to 1 mm. We assumed the density of the dust material to be 1,000 kg m -- 3, the
scattering efficiency for radiation pressure to be 1.0 and a size-dependent dust ejection speed
(at the boundary of the collisional zone) ve ~ a−0.5. The dust emission rate is assumed to be
inversely proportional to the square of the heliocentric distance with the earliest particles
ejected 70 d before the observation. Under these assumptions, we investigated three different
ejection patterns: (1) isotropic; (2) into the subsolar hemisphere with the emission rate
proportional to the cosine of the solar zenith distance; and (3) into a sunward conical jet with
an opening angle of 30° and a constant intensity. The model was fitted to our highest signal-
to-noise ratio r′-band image created from the Gemini data (Fig. 2). The best fit was obtained
for the hemispheric pattern (Supplementary Fig. 3), though the fit of the isotropic emission is
nearly as good. As a result of this procedure, we retrieved a power-law particle size-
distribution exponent of -- 3.7 ± 1.8 and the ejection speed for particles of a given size, equal to
44 ± 14 m s−1 for β = 1 (a = 1.2 μm).
Acknowledgements
Based in part on observations obtained at the Gemini Observatory, which is operated by the
Association of Universities for Research in Astronomy, Inc., under a cooperative agreement
with the NSF on behalf of the Gemini partnership: the National Science Foundation (United
States), the National Research Council (Canada), CONICYT (Chile), Ministerio de Ciencia,
Tecnología e Innovación Productiva (Argentina), and Ministério da Ciência, Tecnologia e
Inovação (Brazil). The William Herschel Telescope is operated on the island of La Palma by
the Isaac Newton Group of Telescopes in the Spanish Observatorio del Roque de los
Muchachos of the Instituto de Astrofísica de Canarias. We thank J. Blakeslee for rapid
evaluation and approval of our Gemini North director's discretionary time request and
P. Jonker for sharing time on the William Herschel Telescope. We also thank the staff of both
observatories for assistance and vital contributions to making these observations possible.
M.D. and P.G. are grateful for support from the National Science Centre of Poland through
SONATA BIS grant no. 2016/22/E/ST9/00109 and Polish Ministry of Science and Higher
Education grant no. DIR/WK/2018/12. G.C. acknowledges support from European Research
Council Consolidator Grant 647208. I.P.-M. acknowledges funding from the Netherlands
Research School for Astronomy (grant no. NOVA5-NW3-10.3.5.14).
References
1. Sekanina, Z. A probability of encounter with interstellar comets and the likelihood of their
existence. Icarus 27, 123 -- 133 (1976).
2. Engelhardt, T. et al. An observational upper limit on the interstellar number density of
asteroids and comets. Astron. J. 153, 133 (2017).
3. Williams, G. MPEC 2017-U181: Comet C/2017 U1 (PanStarrs). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K17/K17UI1.html (2017).
4. Meech, K. J. et al. A brief visit from a red and extremely elongated interstellar asteroid.
Nature 552, 378 -- 381 (2017).
5. Drahus, M. et al. Tumbling motion of 1I/'Oumuamua and its implications for the body's
distant past. Nat. Astron. 2, 407 -- 412 (2018).
6. Micheli, M. et al. Non-gravitational acceleration in the trajectory of 1I/2017 U1
('Oumuamua). Nature 559, 223 -- 226 (2018).
7. The 'Oumuamua ISSI Team. The natural history of 'Oumuamua. Nat. Astron. 3, 594 -- 602
(2019).
8. Bialy, S. & Loeb, A. Could solar radiation pressure explain 'Oumuamua's peculiar
acceleration? Astrophys. J. Lett. 868, L1 (2018).
9. Mamajek, E. Kinematics of the interstellar vagabond 1I/'Oumuamua (A/2017 U1). Res.
Notes AAS 1, 21 (2017).
10. Guzik, P. et al. Interstellar comet gb00234. Astronomer's Telegram 13100 (2019);
http://www.astronomerstelegram.org/?read=13100
11. MPEC 2019-R106: Comet C/2019 Q4 (Borisov). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K19/K19RA6.html (2019).
12. MPEC 2019-R113: Comet C/2019 Q4 (Borisov). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K19/K19RB3.html (2019).
13. MPEC 2019-S03: Comet C/2019 Q4 (Borisov). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K19/K19S03.html (2019).
14. MPEC 2019-S09: Comet C/2019 Q4 (Borisov). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K19/K19S09.html (2019).
15. MPEC 2019-S25: Comet C/2019 Q4 (Borisov). IAU Minor Planet Center
https://www.minorplanetcenter.net/mpec/K19/K19S25.html (2019).
16. Sekanina, Z. & Kracht, R., Preperihelion outbursts and disintegration of comet C/2017 S3
(Pan-STARRS). Preprint at https://arxiv.org/abs/1812.07054 (2018).
17. Fukugita, M. et al. The Sloan Digital Sky Survey photometric system. Astron. J. 111,
1748 -- 1756 (1996).
18. Abazajian, K. N. et al. The seventh data release of the Sloan Digital Sky Survey.
Astrophys. J. Suppl. Ser. 182, 543 -- 558 (2009).
19. Holmberg, J., Flynn, C. & Portinari, L. The colours of the Sun. Mon. Not. R. Astron. Soc.
367, 449 -- 453 (2006).
20. De León, J. et al. A physical characterization of comet C/2019 Q4 (Borisov) with OSIRIS
at the 10.4 m GTC. Res. Notes AAS 3, 9 (2019).
21. Waniak, W., Zoła, S. & Krzesiński, J. Dust emission for comets Shoemaker-Levy 1991a1
and McNaught-Russell 1993v. Icarus 136, 280 -- 297 (1998).
22. Boe, B. et al. The orbit and size-frequency distribution of long period comets observed by
Pan-STARRS1. Icarus 333, 252 -- 272 (2019).
23. Snodgrass, C. et al. The size distribution of Jupiter family comet nuclei. Mon. Not. R.
Astron. Soc. 414, 458 -- 469 (2011).
24. Drahus, M. Rotational disruption of comets with parabolic orbits. In DPS Meet. 46,
200.04 (AAS, 2014).
25. Jewitt, D. Color systematics of comets and related bodies. Astron. J. 150, 201 (2015).
26. Solontoi, M. et al. Ensemble properties of comets in the Sloan Digital Sky Survey. Icarus
218, 571 -- 584 (2012).
27. Jewitt, D. The active Centaurs. Astron. J. 137, 4296 -- 4312 (2009).
28. Fulle, M. in Comets II (eds Festou, M. et al.) 565 -- 575 (Univ. Arizona Press, 2004).
29. Fulle, M., Mikuž, H. & Bosio, S. Dust environment of comet Hyakutake 1996B2. Astron.
Astrophys. 324, 1197 -- 1205 (1997).
30. Jewitt, D. et al. Interstellar interloper 1I/2017 U1: observations from the NOT and WIYN
telescopes. Astrophys. J. Lett. 850, L36 (2017).
31. Do, A., Tucker, M. A. & Tonry, J. Interstellar interlopers: number density and origin of
'Oumuamua-like objects. Astrophys. J. Lett. 855, L10 (2018).
32. Moro-Martín, A., Turner, E. L. & Loeb, A. Will the Large Synoptic Survey Telescope
detect extra-solar planetesimals entering the Solar System? Astrophys. J. 704, 733 -- 742
(2009).
33. Marsden, B. G., Sekanina, Z. & Yeomans, D. K. Comets and nongravitational forces. V.
Astron. J. 78, 211 -- 225 (1973).
34. Cowan, J. J. & A'Hearn, M. F. Vaporization of comet nuclei: light curves and life times.
Moon Planets 21, 155 -- 171 (1979).
35. Jorda, L., Crovisier, J. & Green, D. W. E. The correlation between visual magnitudes and
water production rates. In Asteroids, Comets, Meteors 2008 1405, 8046 (LPI, 2008).
36. Rodgers, C. T. et al. Improved u′g′r′i′z′ to UBVRCIC transformation equations for main-
sequence stars. Astron. J. 132, 989 -- 993 (2006).
37. A'Hearn, M. F. et al. Comet Bowell 1980b. Astron. J. 89, 579 -- 591 (1984).
38. Weiler, M. et al. The dust activity of comet C/1995 O1 (Hale-Bopp) between 3 AU and 13
AU from the Sun. Astron. Astrophys. 403, 313 -- 322 (2003).
39. Campins H. & Fernández Y. Observational constraints on surface characteristics of comet
nuclei. Earth Moon Planets 89, 117 -- 134 (2002).
40. Sekanina, Z. in Comets (ed. Wilkening, L. L.) 251 -- 287 (Univ. Arizona Press, 1982).
41. Samarasinha, N. H. et al. In ESA Proc. 20th ESLAB Symposium on the Exploration of
Halley's Comet Vol. 1, ESA SP-250, 487 -- 491 (ESA, 1986).
42. Jewitt, D. Cometary rotation: an overview. Earth Moon Planets 79, 35 -- 53 (1997).
43. Drahus, M. & Waniak, W. Non-constant rotation period of comet C/2001 K5 (LINEAR).
Icarus 185, 544 -- 557 (2006).
44. Davidsson, B. J. R. Tidal splitting and rotational breakup of solid biaxial ellipsoids. Icarus
149, 375 -- 383 (2001).
45. SDSS photometric equations. SDSS
http://classic.sdss.org/dr7/algorithms/jeg_photometric_eq_dr1.html (2005).
Author contributions
K.R. and P.G. developed Interstellar Crusher. P.G. and M.D. designed the observations, wrote
the telescope time proposal, performed photometry, estimated the size of the nucleus and
wrote the paper. P.G. computed the orbit. M.D. compared the colour to Solar System comets
and estimated the probability of rotational disruption. W.W. reduced raw images and
performed Monte Carlo dust modelling. G.C. and I.P.-M. obtained data at the William
Herschel Telescope.
Data availability
The ACAM data are available from the corresponding authors upon reasonable request. The
GMOS-N raw data will be available in the Gemini Observatory archive at
https://archive.gemini.edu after the expiration of the 12 month proprietary period.
Supplementary Information
Supplementary Fig. 1 Median-stacked images of 2I/Borisov from the William Herschel
Telescope. The top panels show the images obtained on 10 September 2019 at 05:38 UT and
the bottom panels show the images obtained on 13 September 2019 at 05:47 UT (observation
mid-points). The images in the left panels were obtained in the g′ band and the images in the
right panels were obtained in the r′ band. All panels subtend 1.0 × 1.0 arcmin and are scaled
logarithmically. Arrows show the directions of north (N) and east (E), the projected antisolar
vector ( -- ⊙) and the negative of the orbital velocity vector ( -- V). The pixel scale is 0.25 arcsec
px−1 and the seeing was 1.3 arcsec on both nights. The data from 13 September 2019 UT are
affected by stronger twilight and moonlight, and possibly also by a thin cirrus cloud, resulting
in the reduced apparent coma size.
Supplementary Fig. 2 Visibility prospects of 2I/Borisov for years 2019 -- 2020. Top panel
presents basic geometric circumstances and bottom panel shows r′-band magnitude in a
5,000-km radius aperture extrapolated from our Gemini North and WHT photometric data of
10 September 2019 UT. The magnitude is forecasted using three different activity models, in
which the heliocentric brightness of the coma follows the sublimation curves of H2O, CO2 and
CO computed in a standard manner with an assumed Bond albedo of 4% and emissivity of
100%. The actual observed magnitude additionally accounts for the nucleus brightness (we
assumed a 4% geometric albedo and a 0.04 mag deg−1 phase function), phase function of the
coma (0.02 mag deg−1), and geocentric term (inverse square law for brightness).
Supplementary Fig. 3 Synthetic image of 2I/Borisov computed with our Monte-Carlo
code. The model was fitted to the r′-band image obtained at Gemini North on 10 September
2019 at 14:57 UT (Fig. 2, right panel) and has the same orientation and scale. See Methods
for details of the computation.
|
1702.00013 | 1 | 1702 | 2017-01-31T19:00:02 | Four Sub-Saturns with Dissimilar Densities: Windows into Planetary Cores and Envelopes | [
"astro-ph.EP"
] | We present results from a Keck/HIRES radial velocity campaign to study four sub-Saturn-sized planets, K2-27b, K2-32b, K2-39b, and K2-108b, with the goal of understanding their masses, orbits, and heavy element enrichment. The planets have similar sizes $(R_P = 4.5-5.5~R_E)$, but have dissimilar masses $(M_P = 16-60~M_E)$, implying a diversity in their core and envelope masses. K2-32b is the least massive $(M_P = 16.5 \pm 2.7~M_E)$ and orbits in close proximity to two sub-Neptunes near a 3:2:1 period commensurability. K2-27b and K2-39b are significantly more massive at $M_P = 30.9 \pm 4.6~M_E$ and $M_P = 39.8 \pm 4.4~M_E$, respectively, and show no signs of additional planets. K2-108b is the most massive at $M_P = 59.4 \pm 4.4~M_E$, implying a large reservoir of heavy elements of about $\approx50~M_E$. Sub-Saturns as a population have a large diversity in planet mass at a given size. They exhibit remarkably little correlation between mass and size; sub-Saturns range from $\approx 6-60~M_E$, regardless of size. We find a strong correlation between planet mass and host star metallicity, suggesting that metal-rich disks form more massive planet cores. The most massive sub-Saturns tend to lack detected companions and have moderately eccentric orbits, perhaps as a result of a previous epoch of dynamical instability. Finally, we observe only a weak correlation between the planet envelope fraction and present-day equilibrium temperature, suggesting that photo-evaporation does not play a dominant role in determining the amount of gas sub-Saturns accrete from their protoplanetary disks. | astro-ph.EP | astro-ph | Draft version February 2, 2017
Preprint typeset using LATEX style AASTeX6 v. 1.0
FOUR SUB-SATURNS WITH DISSIMILAR DENSITIES: WINDOWS INTO PLANETARY CORES AND
ENVELOPES
Erik A. Petigura1,2,13, Evan Sinukoff3,14, Eric Lopez4, Ian J. M. Crossfield5,15, Andrew W. Howard1, John
M. Brewer6, Benjamin J. Fulton1,3, Howard T. Isaacson7, David R. Ciardi8, Steve B. Howell9, Mark E.
Everett10, Elliott P. Horch11, Lea Hirsch7, Lauren M. Weiss12,16, and Joshua E. Schlieder8
7
1
0
2
n
a
J
1
3
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
3
1
0
0
0
.
2
0
7
1
:
v
i
X
r
a
1California Institute of Techonology
[email protected]
3Institute for Astronomy, University of Hawai'i at M¯anoa
4Institute for Astronomy, University of Edinburgh
5University of California, Santa Cruz
6Yale University
7University of California, Berkeley
8IPAC-NExScI, California Institute of Technology
9NASA Ames Research Center
10National Optical Astronomy Observatory
11Department of Physics, Southern Connecticut State University
12Institut de Recherche sur les Exoplanètes, Université de Montréal
13Hubble Fellow
14NSERC Fellow
15NASA Sagan Fellow
16Trottier Fellow
ABSTRACT
We present results from a Keck/HIRES radial velocity campaign to study four sub-Saturn-sized plan-
ets, K2-27b, K2-32b, K2-39b, and K2-108b, with the goal of understanding their masses, orbits, and
heavy element enrichment. The planets have similar sizes (RP = 4.5–5.5 R⊕), but have dissimilar
masses (MP = 16–60 M⊕), implying a diversity in their core and envelope masses. K2-32b is the
least massive (MP = 16.5 ± 2.7 M⊕) and orbits in close proximity to two sub-Neptunes near a 3:2:1
period commensurability. K2-27b and K2-39b are significantly more massive at MP = 30.9 ± 4.6 M⊕
and MP = 39.8 ± 4.4 M⊕, respectively, and show no signs of additional planets. K2-108b is the most
massive at MP = 59.4 ± 4.4 M⊕, implying a large reservoir of heavy elements of about ≈ 50 M⊕.
Sub-Saturns as a population have a large diversity in planet mass at a given size. They exhibit
remarkably little correlation between mass and size; sub-Saturns range from ≈6–60 M⊕, regardless
of size. We find a strong correlation between planet mass and host star metallicity, suggesting that
metal-rich disks form more massive planet cores. The most massive sub-Saturns tend to lack detected
companions and have moderately eccentric orbits, perhaps as a result of a previous epoch of dynami-
cal instability. Finally, we observe only a weak correlation between the planet envelope fraction and
present-day equilibrium temperature, suggesting that photo-evaporation does not play a dominant
role in determining the amount of gas sub-Saturns accrete from their protoplanetary disks.
Keywords: editorials, notices - miscellaneous - catalogs - surveys
1. INTRODUCTION
The Solar System contains four terrestrial planets, two
ice giants, and two gas giants on nearly circular and co-
planar orbits. Notably, the Solar System lacks several
broad classes of planets:
it contains no planets having
sizes between Earth and Neptune (1.0–3.9 R⊕) or be-
tween Uranus and Saturn (4.0–9.4 R⊕), and no planets
that orbit closer than Mercury (0.39 AU). A longstand-
ing question is whether the Solar System is representa-
tive of planetary systems around other stars, or if it is
one particular realization of a set of physical processes
that produce a diversity of outcomes.
The study of extrasolar planets offers a path to ad-
2
dress this question. Studies of planet occurrence from
the prime Kepler mission (Borucki et al. 2010b) revealed
that our Solar System is atypical in a few key ways: the
majority of stars have at least one planet interior to Mer-
cury's orbit and the most common size of planet is in the
range 1–3 R⊕, sizes not represented in our Solar System
(Howard et al. 2012; Fressin et al. 2013; Petigura et al.
2013). The occurrence of planets rises rapidly below
RP = 3.0 R⊕, indicating an important size scale in the
formation of planet cores and envelopes. Constraining
the bulk composition of these sub-Neptunes has been
the focus of intensive radial velocity (RV) campaigns
that revealed that most planets larger than ≈ 1.6 R⊕
have significant gaseous envelopes (Marcy et al. 2014;
Weiss & Marcy 2014; Rogers 2015).
In this paper, we focus on another size class of plan-
ets absent in the Solar System, sub-Saturns, which we
define as planets having sizes between 4.0–8.0 R⊕. Sub-
Saturns offer a superb laboratory to study planet forma-
tion history and compositions. Their large sizes require
significant envelopes of H/He. For sub-Saturns, H/He
comprise such a large component of the planet volume
that the planets can be modeled as a high-density core
with a thick H/He envelope. Measurements of the core
mass fraction are simplified because details in the com-
position of the core have little effect on the measured
planet size (Lopez & Fortney 2014; Petigura et al. 2016).
While close-in (< 1 AU) gaseous planets are thought
to form via core accretion (Pollack et al. 1996; Boden-
heimer et al. 2000; Hubickyj et al. 2005; Mordasini et al.
2008) there are major uncertainties regarding how plan-
ets acquire (and lose) mass and the extent to which their
orbits evolve with time. For example, Jupiter is thought
to have taken ∼3 Myr to accrete enough gas before trig-
gering runaway accretion (Hubickyj et al. 2005) while
gas disks around other stars are observed to dissipate
after 1–10 Myr (Mamajek 2009). That these two pro-
cesses have similar timescales may explain why the oc-
currence of Jovians within 20 AU is ≈ 20% rather than
≈ 100% (Cumming et al. 2008).
While sub-Saturns are similar to Jovians given their
H/He envelopes, they often have much lower masses.
For example, Kepler-79d is 7.2 R⊕ and only 6.0 M⊕
(Jontof-Hutter et al. 2014). For sub-Saturns, the run-
away accretion of gas invoked to explain Jupiter's mas-
sive envelope seems to have not occurred. Low-density
sub-Saturns have inspired alternative gas accretion sce-
narios, such as accretion in a gas-depleted disk (e.g. Lee
& Chiang 2015a).
Here, we present RV measurements of
four Sub-
Saturns, K2-27b, K2-32b, K2-39b, and K2-108b, taken
as part of a program to expand the sample of sub-
Saturns with well-measured masses and radii. These
planets were observed by the Kepler Space Telescope
(Borucki et al. 2010b) operating during its K2 mis-
sion, where the telescope observes a new field in the
ecliptic plane every ≈ 3 months (Howell et al. 2014).
Sections 2-4 present the radial velocity measurements,
stellar characterization, and modeling needed to extract
planet mass, radius, and orbital eccentricity. For these
planets, we achieve mass measurements of 16% or better
and density measurements to 33% or better.
In Section 5, we place the four planets in the context
of other sub-Saturns. We find a large diversity in planet
masses in the sub-Saturn size range with little corre-
lation with planet size. Sub-Saturns range from ≈6–
60 M⊕, regardless of their size. We find a strong corre-
lation between stellar metallicity and planet mass, sug-
gesting metal-rich disks likely form more massive planet
cores. We also observe that planet mass seems to be in-
versely correlated with the presence of additional plan-
ets, which could be the result of a period of large-scale
dynamical instabilities resulting in mergers or scatter-
ing on to high inclination orbits. We apply the interior
structure models of (Lopez & Fortney 2014) to deter-
mine the fraction of planet mass in H/He and heavy el-
ement. For sub-Saturns, we see only a weak dependence
of the envelope fraction on equilibrium temperature, in-
dicating that photo-evaporation does not likely play a
major role in sculpting the final sizes of sub-Saturns.
We offer some concluding thoughts in Section 6.
2. RADIAL VELOCITY OBSERVATIONS AND
ANALYSIS
Here, we describe our overall RV observational cam-
paign and our analysis methodology. Details on individ-
ual systems are given in Section 4. We observed K2-27,
K2-32, K2-39, and K2-108 using the High Resolution
Echelle Spectrometer (HIRES; Vogt et al. 1994) on the
10 m Keck Telescope I. We collected spectra through an
iodine cell mounted directly in front of the spectrometer
slit. The iodine cell imprints a dense forest of absorption
lines which serve as a wavelength reference. We used an
exposure meter to achieve a consistent signal to noise
level for each program star, which ranged from 100 to
130 per reduced pixel on blaze near 550 nm. We also
obtained a "template" spectrum without iodine.
RVs were determined using standard procedures of the
California Planet Search (CPS; Howard et al. 2010) in-
cluding forward modeling of the stellar and iodine spec-
tra convolved with the instrumental response (Marcy &
Butler 1992; Valenti et al. 1995). The RVs are tabulated
in Table 1. We also list the measurement uncertainty of
each RV point, which ranges from 1.5 to 2.0 m s−1 and
is derived from the uncertainty on the mean RV of the
∼700 spectral chunks used in the RV pipeline.
We analyzed the RV time-series using the publicly-
available RV-fitting package radvel (Fulton & Petigura,
in prep.)1 When modelling the RVs, we adopt the like-
lihood, L, of Howard et al. (2014):
(cid:34)
(cid:88)
i
lnL = − 1
2
(vi − vm,i)2
i + σ2
σ2
jit
+ ln 2π(cid:0)σ2
i + σ2
jit
,
(cid:1)(cid:35)
where vi is the i'th RV measured at time ti, σi is the
corresponding uncertainty, vm is the Keplerian model
velocity at ti, and σjit or "jitter" accounts for additional
RV variability due to stellar and instrumental noise and
is included in our models as a free parameter.
When modeling the RVs, we first consider circular Ke-
plerians with no additional acceleration term, γ. Here,
σjit and an average RV offset, γ, are allowed to float
as free parameters. We then allow for more complicated
models if they are motivated by the data. We consider
models where γ is allowed to float and models where
eccentricity, e, and longitude of periastron, ω(cid:63), are al-
lowed to vary.2 More complex models will naturally lead
to higher likelihoods at the expense of additional free
parameters. To assess whether a more complex model
is justified, we use the Bayesian Information Criterion
(BIC; Schwarz 1978). Models with smaller BIC, i.e. neg-
ative ∆BIC are preferend.
When available, we incorporated RV measurements
from the literature to augment our HIRES timeseries.
RVs of K2-27, K2-32, and K2-39 have been published
in Van Eylen et al. (2016a), Dai et al. (2016), and Van
Eylen et al. (2016b), respectively. We fit for the offset
and jitter terms independently for different datasets. To
derive uncertainties on the RV parameters we perform
a standard MCMC exploration of the likelihood surface
using the emcee Python package (Goodman & Weare
2010; Foreman-Mackey et al. 2013). In Tables 3–6, we
list orbital and planetary properties assuming both cir-
cular and eccentric orbits. The preferred model is indi-
cated.
Table 1. Radial Velocities
Star
Inst.
Time
BJDTBD
RV
m s−1
−3.30
K2-27 HIRES
2457059.023437
2457187.504940 −37782.19
K2-27 HARPS
K2-27 HARPS-N 2457064.713740 −37785.55
2457045.607900 −38039.00
K2-27 FIES
σ(RV)
m s−1
3.58
2.56
6.56
11.10
Table 1 continued
√
1 https://github.com/California-Planet-Search/radvel
2 During fitting and MCMC modeling, we parametrize e and
ω(cid:63) by
e sin ω(cid:63), as recommended by Eastman et al.
(2013), to guard against the Lucy-Sweeney bias toward non-zero
eccentricities.
e cos ω(cid:63) and
√
Table 1 (continued)
3
Star
Inst.
K2-32 HIRES
K2-32 HARPS
K2-32 PFS
K2-39 HIRES
K2-39 HARPS
K2-39 PFS
K2-39 FIES
Time
BJDTBD
RV
m s−1
σ(RV)
m s−1
2457179.918605
2457185.606900
2457198.674600
2457245.118029
2457255.714330
2457257.799090
2.24
10.69
−13.95
−2.77
24507.93
0.73
2457235.669620
24557.22
1.81
2.65
2.31
1.84
2.66
1.64
6.99
Note-The radial velocity (RV) measurements used in this work.
We list the HIRES RVs along with other RVs from the literature,
where available. Table 1 is published in its entirety in machine-
readable format. A portion is shown here for guidance regarding
its form and content.
3. STELLAR PROPERTIES
log g, and metallicity,
We measured stellar effective temperature, Teff, sur-
face gravity,
[Fe/H] from our
iodine-free "template" spectra. We followed the method-
ology of Brewer et al. (2016), which used an updated
version of the SME code and has been shown to recover
surface gravities consistent with those from asteroseis-
mology to within 0.05 dex (Brewer et al. 2015). We con-
strained stellar mass, M(cid:63), radius, R(cid:63), and age from Teff,
log g, and [Fe/H] using the isochrones Python pack-
age (Morton 2015) which interpolates among the Dart-
mouth stellar isochrones (Dotter et al. 2008). The stellar
properties are listed alongside other system properties in
Tables 3-6, respectively.
For K2-39, we note some tension between the stel-
lar parameters presented by Van Eylen et al. (2016b)
and those presented here.
Importantly, the log g de-
rived by Van Eylen et al. (2016b) is lower than the
SME value. The Van Eylen et al. (2016b) analysis re-
sulted in a larger inferred stellar radius measurement
that, at R(cid:63) = 3.90+0.30−0.27 R(cid:12),
is 34% larger than the
R(cid:63) = 2.93 ± 0.21 R(cid:12) derived by SME and isochrones.
We present a side-by-side comparison of spectroscopic
parameters in Table 2.
Can the recently released Gaia parallaxes be used to
resolve these different estimates of stellar radius? K2-
39 is listed in the Tycho catalog as TYC-5811-835-1.
Gaia recently released parallaxes for most stars in the
2 million-star Tycho catalog. K2-39 has a parallax of
3.35 ± 0.86 mas. Given the apparent K-band magni-
tude of 8.516 ± 0.024 from 2MASS (Cutri et al. 2003),
we calculated the parallax implied by both sets of spec-
troscopic parameters. We used K-band since it is less
sensitive to the unknown amount of extinction between
Earth and K2-39. When we used the SME parameters we
4
found parallax of 3.50+0.24−0.20 mas, in close agreement with
the Gaia parallax. The larger radius of Van Eylen et al.
(2016b) necessitates a more distant star to produce the
observed K mag, and thus yields a smaller expected par-
allax of 2.62 ± 0.20 mas. While this parallax is smaller
than the most likely Gaia value, it is still consistent at
the 1σ level. Given that Gaia Data Release 1 provided a
4σ measurement of K2-39's parallax, it is insufficient to
distinguish between the two sets of parameters. Future
Gaia releases will provide important constraints on the
physical properties of K2-39.
Van Eylen et al. (2016b) also noted that the non-
detection of asteroseismic modes places a lower limit of
log g ≥ 3.50 dex. Taken together, the non-dection of
asteroseismic modes and the Gaia parallax both suggest
that K2-39 is smaller than reported in Van Eylen et al.
(2016b). Given that SME has been extensively tested
against asteroseismology we adopt the SME parameters
hereafter.
Table 2. K2-39 Stellar Parameters
Teff (K)
log g (dex)
[Fe/H] (dex)
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
Parallax (mas)a
This work
4912 ± 60
3.58 ± 0.05
0.43 ± 0.04
1.192+0.085
−0.070
2.93 ± 0.21
3.50+0.24−0.20
V16
4881 ± 20
3.44 ± 0.07
0.32 ± 0.04
1.35+0.04−0.03
3.90+0.30−0.27
2.62 ± 0.20
Note-Comparison of the stellar properties from
this work (based on SME, Brewer et al. 2015), and
from Van Eylen et al. (2016b). M(cid:63) and R(cid:63) in
the Van Eylen et al. (2016b) column were derived
using the isochrones package which interpolates
between Dartmouth isochrones, as opposed to
Van Eylen et al. (2016b) who used Yonsei-Yale
isochrones. However, Van Eylen et al. (2016b)
gives R(cid:63) = 3.88+0.48−0.42 R(cid:12), which is consistent with
the Dartmouth model. We adopt the SME value
given that the code has been extensively tested
against asteroseismology.
aImplied parallax from based on spectroscopic
properties, isochrone modeling, and apparent K-
band magnitude.
4. INDIVIDUAL TARGETS
4.1. K2-27
K2-27 hosts a single transiting sub-Saturn, K2-27b,
with P = 6.77 d that was first confirmed in Van Eylen
et al. (2016a) using RVs. We obtained 15 spectra with
HIRES of K2-27 between 2015-02-05 and 2016-07-17.
Van Eylen et al. (2016a) observed this star with HARPS,
HARPS-N, and FIES. We included 6 and 19 measure-
ments from HARPS and HARPS-N, respectively in our
RV analysis. We did not include the 6 FIES measure-
ments because their uncertainties are much larger (≈
30 m s−1), compared to uncertainties from the HIRES,
HARPS, and HARPS-N spectra (≈ 4 − 5 m s−1) and
hence add little additional information to constrain the
RV fits while increasing model complexity.
We first modeled the combined RVs assuming circular
orbits and no additional acceleration term, γ. We then
allowed γ to vary, but found that these models produced
a negligible improvement in the BIC (∆BIC = −1). We
then allowed for eccentricity to float and found a non-
zero eccentricity of e = 0.251 ± 0.088. Compared to the
circular models, the eccentric fits were strongly favored
(∆BIC = −14). We verified that the eccentricity is de-
tected in both HIRES and HARPS+HARPS-N datasets
by fitting these subsets independently. These datasets
yield consistent and non-zero eccentricities. We adopted
the parameters from the eccentric model as the pre-
ferred parameters. The most probable eccentric model
is shown in Figure 1. The K2-27 system parameters
are summarized in Table 3. We measured a Doppler
semi-amplitude of K = 11.8 ± 1.8 m s−1, which is con-
sistent with K = 10.8±2.7 m s−1 reported by Van Eylen
et al. (2016a), but with smaller uncertainties due to
our additional measurements. We measured a mass of
MP = 30.9±4.6 M⊕ and a density of 1.87±0.41 g cm−3.
As we discuss in Section 5, K2-27b has a relatively high
mass for its size, implying a large core-mass.
4.2. K2-32
K2-32 hosts three planets, K2-32b, K2-32c, and K2-
32d, having orbital periods of P = 8.99 d, 20.66 d,
and 31.7 d, respectively, which are near the 3:2:1 period
commensurability. The planets were first confirmed in
Sinukoff et al. (2016) using multiplicity arguments (Lis-
sauer et al. 2012). We obtained 31 spectra of K2-32
with HIRES between 2015-06-06 and 2016-08-20. Dai
et al. (2016) obtained 43 spectra with HARPS and 6
with PFS, which we included in our RV analysis. We
first modeled the combined HIRES, HARPS, and PFS
RVs assuming circular orbits and no additional acceler-
ation term, γ. When allowing γ to float, we found that
γ = 2.3 ± 1.8 m s−1 yr−1, consistent with zero at the
2σ level. This differs from γ = 34.0+9.9−9.7 m s−1 yr−1 re-
ported by Dai et al. (2016). All but 5 of the RVs from
Dai et al. (2016) were collected within a 25-day window,
and thus do not provide much leverage on γ. The posi-
tive γ measured by Dai et al. (2016) is driven by two RV
measurements that fall above our best fit curve. With
our combined dataset spanning two observing seasons,
we place tighter limits on γ and on the presence of long
period companions having P (cid:38) 2 years. Nonetheless,
the BIC preferred fixing γ at 0 m s−1 yr−1 (∆BIC = 1).
Next, we considered eccentric models. In the circular
5
ensemble of mass measurements from TTVs and RVs,
K2-32b is of intermediate density (see Section 5).
Figure 1. Single Keplerian model of K2-27 radial velocities (RVs), allowing for eccentricity (see Section 4.1). a) Time series of
RVs from HIRES along with HARPS and HARPS-N published in Van Eylen et al. (2016a). During the fitting, we allowed for
an arbitrary offset between the three instruments to float as a free parameter. The blue line shows the most probable Keplerian
model. b) Residuals to the most probable Keplerian model. c) Phase-folded RVs and the most probable Keplerian.
models, Kb = 1.6± 1.0 m s−1 and Kc = 2.3± 1.1 m s−1,
respectively. Given that the reflex motion due to K2-
32c and K2-32d were only detected at the 2σ level, we
cannot place meaningful constraints on their eccentric-
ities. Therefore, we fixed their eccentricities at zero in
our fits and allowed the eccentricity of K2-32b to float.
We found that the eccentricity of K2-32b was consistent
with zero, and placed an upper limit of eb < 0.23 (95%
conf.). We therefore adopted the circular model for the
K2-32 system parameters. The physical properties of
the K2-32 system along with our circular and eccentric
models are listed Table 4. The best-fit circular model is
shown in Figure 2.
For K2-32b, we measured MP = 16.5± 2.7 M⊕, which
is lower than, but within the 1σ confidence interval of
MP = 21.1 ± 5.9, measured by Dai et al. (2016). We
measure a density of ρ = 0.67±0.16 g cm−3, which is low
compared to other sub-Saturns of similar size with RV
mass measurements. However, when compared to the
Figure 2 shows that the Dai et al. (2016) measure-
ments were not well-sampled during the quadrature
times of K2-32c and K2-32d, leading to poor constraints
on their Doppler semi-amplitudes with significant co-
variance between Kc and Kd due the 3:2 period ratios
of the planets. Our combined dataset has more uniform
sampling in phase with minimal covariance between Kc
and Kd.
We achieve marginal detections of K2-32c and K2-32d,
which have masses of 6.2 ± 3.9 M⊕ and 10.3 ± 4.7 M⊕,
respectively. Because K2-32c is not quite a 2σ detec-
tion, we conservatively report an upper limit of MP
< 12.1 M⊕ (95% conf.), to be conservative. Contin-
ued Doppler monitoring of K2-32 is necessary to place
tighter constraints on the masses and orbits of K2-32c
and K2-32d. While such observations are challeng-
ing with current instruments given the faint host star
−30−20−10010203040RV [m s-1]a)HARPSHARPS-NHIRES2015.22015.42015.62015.82016.02016.22016.471007200730074007500BJDTDB - 2450000−20020Residualsb)−0.4−0.20.00.20.4Phase−30−20−100102030RV [m s-1]c)Pb = 6.77 daysKb = 11.87 m s-1eb = 0.24 6
Table 3. System parameters of K2-27
Value Ref.
Stellar parameters
Identifier
Teff (K)
log g (dex)
[Fe/H] (dex)
v sin i (km s−1)
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
age (Gyr)
Apparent V (mag)
EPIC-201546283
5248 ± 60
4.48 ± 0.05
0.13 ± 0.04
2.3
0.866+0.029
−0.023
0.885 ± 0.043
10.3+3.3−5.2
12.64 ± 0.02
planet b
6.771315 ± 0.000085
1979.84484 ± 0.00057
4.48 ± 0.23
0.06702 ± 0.00071
116+16−12
902 ± 28
A
A
A
A
A
A
A
A
B
B
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
Transit model
P (days)
T0 (BJD-2454833)
RP (R⊕)
a (AU)
Sinc (S⊕)
Teq (K)
Circular RV model
K (m s−1)
γHIRES (m s−1)
γHARPS (m s−1)
γHARPS−N (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,HARPS−N (m s−1)
MP (M⊕)
ρ (g cm−3)
Eccentric RV model (adopted)
K (m s−1)
e
γHIRES (m s−1)
γHARPS (m s−1)
γHARPS−N (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,HARPS−N (m s−1)
MP (M⊕)
ρ (g cm−3)
9.9 ± 2.0
−2.7 ± 2.0
−0.3 ± 3.8
6.4 ± 2.2
0 (fixed)
6.1+2.0−1.4
6.9+5.9−3.2
4.7 ± 2.4
26.7 ± 5.3
1.61 ± 0.42
11.8 ± 1.8
0.251 ± 0.088
−2.2 ± 1.6
0.2 ± 2.7
6.4 ± 2.2
0 (fixed)
4.4+1.8−1.2
4.1+4.8−2.6
4.8+2.6−2.1
30.9 ± 4.6
1.87 ± 0.41
Note-A: This work; B: Crossfield et al. (2016)
(V = 12.3 mag), the dynamical richness of this system
makes these observations worthwhile.
4.3. K2-39
K2-39 hosts a single transiting planet, K2-39b, with
P = 4.60 d, which was first confirmed by Van Eylen
et al. (2016b). When fitting the photometry, Van Eylen
et al. (2016b) and Crossfield et al. (2016) arrived at dif-
ferent values of the planet-to-star radius ratio, RP /R(cid:63),
of 1.93 ± 0.1% and 2.52 ± 0.27%, respectively. The
Crossfield et al. (2016) solution favored a grazing im-
pact parameter of b = 1.10+0.07−0.09 and is responsible for
the larger inferred radius ratio. Motivated by this dis-
crepancy we re-examined the K2-39 photometry pro-
duced by the k2phot pipeline3 and two other publicly
available light curves produced by the k2sff and k2sc
pipelines (Vanderburg & Johnson 2014; Aigrain et al.
2015). For each light curve, we masked out the in-transit
points and modeled the out-of-transit photometry with
a Gaussian Process (Rasmussen & Williams 2005) using
a squared-exponential kernel with a correlation length
of 2 days. We then performed a standard MCMC explo-
ration of the likelihood surface using the batman (Krei-
dberg 2015) and emcee Python packages to map out pa-
rameter uncertainties and covariances. Figure 3 summa-
rizes the results of these different fits. The k2phot, k2sc,
and k2sff light curves yielded RP /R(cid:63) of 1.85+0.15−0.05%,
1.76+0.15−0.06%, and 1.66+0.12−0.05%, respectively. We also note
the significant correlation between RP /R(cid:63) and b at large
values of b, which is responsible for the rather asymmet-
ric uncertainties. Because RP /R(cid:63) differs by 2–3σ among
the different reductions, we combined the three sets of
MCMC chains and adopt RP /R(cid:63) = 1.79 ± 0.13% as an
intermediate value of the planet-to-star radius ratio with
more conservative errors.
We obtained 42 spectra with HIRES of K2-39 between
2015-08-10 and 2016-08-21. Van Eylen et al. (2016b) ob-
tained 7 spectra with HARPS, 6 with PFS, and 17 with
FIES, which we included in our analysis. In contrast to
our K2-27 analysis (Section 4.1), we included FIES RVs
because of the larger number of measurements (17 as
opposed to 6 for K2-27) and because they are the only
non-HIRES dataset with sufficient time baseline to re-
solve a long timescale activity signal, which we discuss
below.
The RVs exhibited large amplitude variability that
was not associated with K2-39b that motivated searches
for additional non-transiting planets. Figure 4 shows
three successive Keplerian searches in the K2-39 RVs
using a modified version of the Two-Dimensional Ke-
plerian Lomb-Scargle (2DKLS) periodogram (O'Toole
et al. 2009; Howard & Fulton 2016). We observed a
peak in the periodogram at P = 4.6 d, which corre-
sponds to the period of K2-39b. When we measured the
change in χ2 (periodogram power) between a 1-planet
fit and a 2-planet fit, then between a 2 planet model
and a 3-planet model (lower two panels in Figure 4) we
found several peaks that fall above the 10% empirical
false alarm threshold (eFAP, Howard & Fulton 2016).
However, inspection of the S-values (Isaacson & Fischer
2010, see Figure 4), which can correlate with stellar ac-
tivity that can lead to spurious Doppler signals, showed
significant long-term variability that is associated with
the RV peak seen at ≈330 d and is likely not associated
3 https://github.com/petigura/k2phot
7
Figure 2. Three-planet Keplerian fit to the K2-32 radial velocities (RVs), assuming circular orbits (see Section 4.2). a) Time
series of RVs from HIRES along with HARPS and PFS published in Dai et al. (2016). During the fitting, we allowed for an
arbitrary offset between the three instruments to float as a free parameter. The blue line shows the most probable Keplerian
model. b) Residuals to the most probable Keplerian model. Panels c) through e) show the phase-folded RVs and the most
probable Keplerian with the contributions from the other planets removed. The large red circles show the phase-binned RVs .
−20−100102030RV [m s-1]a)HARPSHIRESPFS2015.62015.82016.02016.22016.42016.672007300740075007600BJDTDB - 2450000−15015Residualsb)−0.4−0.20.00.20.4Phase−15−10−5051015RV [m s-1]c)Pb = 8.99 daysKb = 5.54 m s-1eb = 0.00 −0.4−0.20.00.20.4Phase−10−50510RV [m s-1]d)Pc = 20.66 daysKc = 1.84 m s-1ec = 0.00 −0.4−0.20.00.20.4Phase−10−50510RV [m s-1]e)Pd = 31.72 daysKd = 2.15 m s-1ed = 0.00 8
Stellar parameters
Identifier
Teff (K)
log g (dex)
[Fe/H] (dex)
v sin i (km s−1)
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
age (Gyr)
Apparent V (mag)
EPIC-205071984
5275 ± 60
4.49 ± 0.05
−0.02 ± 0.04
0.7
0.856 ± 0.028
0.845+0.044
−0.035
7.9 ± 4.5
12.31 ± 0.02
planet b
8.99213 ± 0.00016
2076.91832 ± 0.00055
5.13 ± 0.28
0.08036 ± 0.00088
77.7+10.8−8.3
817 ± 25
Transit model
P (days)
T0 (BJD-2454833)
RP (R⊕)
a (AU)
Sinc (S⊕)
Teq (K)
Circular RV model (adopted)
K (m s−1)
γHIRES (m s−1)
γHARPS (m s−1)
γPFS (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,PFS (m s−1)
MP (M⊕)
ρ (g cm−3)
Eccentric RV model
K (m s−1)
e
γHIRES (m s−1)
γHARPS (m s−1)
γPFS (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,PFS (m s−1)
MP (M⊕)
ρ (g cm−3)
1.6 ± 1.0
5.63 ± 0.91
−1.69 ± 0.85
1.07 ± 0.84
−6.7 ± 3.2
0 (fixed)
3.77+0.81−0.65
4.13 ± 0.73
6.5+4.3−2.4
16.5 ± 2.7 < 12.1 (95% conf.)
0.67 ± 0.16
< 2.7 (95% conf.)
1.7 ± 1.0
Unconstrained
5.60 ± 0.93
< 0.23 (95% conf.)
−1.70 ± 0.87
1.15 ± 0.85
−6.7 ± 3.4
0 (fixed)
3.88+0.82−0.65
4.13 ± 0.73
6.6+4.6−2.5
16.5 ± 2.8 < 12.7 (95% conf.)
< 2.5 (95% conf.)
0.58+0.16−0.13
Table 4. K2-32 System Parameters
Value
Ref.
planet c
20.6602 ± 0.0017
2128.4067 ± 0.0032
3.01 ± 0.25
0.1399 ± 0.0015
25.6+3.6−2.7
619 ± 19
planet d
31.7154 ± 0.0022
2070.7901 ± 0.0026
3.43 ± 0.35
0.1862 ± 0.0020
14.5+2.0−1.5
537 ± 16
A
A
A
A
A
A
A
A
B
B
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
2.3 ± 1.1
10.3 ± 4.7
1.38+0.92−0.67
2.4 ± 1.1
Unconstrained
10.9 ± 4.9
1.29+0.86−0.61
Note-A: This work; B: Crossfield et al. (2016). Because the 2σ confidence interval on Kc includes zero, we report upper limits on the
mass and density of K2-32.
with another planet. We modeled out the activity sig-
nal to search for additional non-transiting planets. As a
matter of convenience, we modeled the activity signature
as a Keplerian and removed it from the timeseries. A
subsequent search of the residual RVs revealed no other
significant signals.
We modeled the combined HIRES, HARPS, FIES,
and PFS RVs using two Keplerians: one for K2-39b and
one as a convenient description of the stellar activity.
We first assumed circular orbits and no additional accel-
eration term, γ. We found that models that included γ
were not favored by the BIC and fixed γ = 0 m s−1 yr−1
in subsequent fits. Next, we allowed the eccentricity of
K2-39b to float, and found that this model was preferred
over the circular model (∆BIC = −7). We adopted the
eccentric model, which is shown in Figure 5. The prop-
erties of the K2-39 system are listed in Table 5.
For K2-39b, we found a mass of MP = 39.8± 4.4 M⊕,
which agrees with MP = 50.3+9.7−9.4 M⊕ reported by Van
Eylen et al. (2016b) at the 1σ level. The additional
RVs improved the precision of the mass measurement
by roughly a factor of two. Our derived planet radius
of RP = 5.71 ± 0.63 R⊕ is substantially smaller than
RP = 8.2±1.1 R⊕ reported by Van Eylen et al. (2016b).
This is largely due to the smaller stellar radius (see Sec-
tion 3). Our adopted value of RP /R(cid:63) = 1.79 ± 0.13% is
also smaller than the Van Eylen et al. (2016b) value of
RP /R(cid:63) = 1.93 ± 0.1%. While the difference in RP /R(cid:63)
is not as large in a fractional sense as the difference in
R(cid:63), it also contributes to a smaller derived RP . Thus,
our derived density of ρ = 1.17+0.47−0.32 g cm−3, is signif-
icantly larger than ρ = 0.50+0.29−0.17 g cm−3, reported by
Van Eylen et al. (2016b).
4.4. K2-108
K2-108,
listed as EPIC-211736671 in the Ecliptic
Planet Input Catalog (Huber et al. 2016), is a V =
12.3 mag star observed during K2 Campaign 5. We
identified K2-108 as a likely planet according to our
team's standard methodology, described in detail
in
Crossfield et al. (2016).
In brief, we identified a set
of transits having P = 4.73 d and elevated K2-108
to the status of "planet candidate." We fit the light
curve according to standard procedures and show the
best-fitting model light curve in Figure 6. Follow-up
spectroscopic observations revealed that K2-108 is a
metal-rich ([Fe/H] = 0.33 ± 0.04 dex), slightly-evolved
G star having a radius of R(cid:63) = 1.75 ± 0.14 R(cid:12). The
spectroscopically-determined stellar parameters along
with the results from our light curve fitting are listed
in Table 6.
We obtained 20 spectra with HIRES of K2-108 be-
tween 2015-12-23 and 2016-11-25. We first considered
circular models with no acceleration term, γ. We found
9
Figure 3. Results from MCMC fitting of different photo-
metric reductions of K2-39 light curve, described in Sec-
tion 4.3. Top: different values of the RP /R(cid:63) from the lit-
erature and this work. Significant disagreement between the
Van Eylen et al. (2016b) and Crossfield et al. (2016) values
(V16 and C16, respectively), motivated a reanalysis of pho-
tometry generated by three independent pipelines: k2phot,
k2sc, and k2sff. The histograms show the posterior distri-
butions after an MCMC exploration for different datasets.
The different reductions led to posteriors on RP and b that
differed by ≈ 1− 3σ, perhaps due to different susceptibilities
to correlated noise. Our adopted value combines all three
chains to conservatively represent RP /R(cid:63). Bottom: the 1σ
and 2σ contours for RP /R(cid:63) and the impact parameter, b.
We see the well-known correlation between RP /R(cid:63) and b for
large values of b. The Crossfield et al. (2016) analysis favored
a grazing transit, and thus a larger value of RP /R(cid:63). Given
the disagreement between the various reductions, we com-
bined the three MCMC chains to derive a more conservative
"adopted" value of RP /R(cid:63).
that eccentric models with non-zero γ (shown in Fig-
ure 7) were favored over circular models (∆BIC =
−25). The parameters are summarized in Table 6. At
59.4 ± 4.4 M⊕, K2-108b is remarkably massive for a
5.28 ± 0.54 R⊕ planet, implying a large heavy element
component.
While our RV analysis verified the planetary nature
of the transiting object, we assessed the possibility of
additional stellar companions in the photometric aper-
ture that could appreciably dilute the observed tran-
sit, resulting in an incorrect derived planetary radius.
From the light curve fits, the planet-to-star radius ratio
020040060080010001200Probability DensityC16V16Adoptedk2photk2sck2sff0.0160.0180.0200.0220.0240.0260.028Rp=R⋆0.00.20.40.60.8b10
Table 5. K2-39 planet parameters
Value
Ref.
activity
A
A
A
A
A
A
A
A
B
B
A
A
A
A
A
B
B
A
A
A
A
A
A
A
A
A
A
A
A
B
B
A
A
A
A
A
A
A
A
A
A
A
A
A
Stellar parameters
Identifiers
EPIC-206247743
TYC-5811-835-1
4912 ± 60
3.58 ± 0.05
0.43 ± 0.04
0.1
1.192+0.085
−0.070
2.93 ± 0.21
6.7+1.7−1.3
10.83 ± 0.08
planet b
4.60497 ± 0.00077
2152.4315 ± 0.0058
1.79 ± 0.13
5.71 ± 0.63
0.0574 ± 0.0012
1356 ± 175
1670 ± 54
Teff (K)
log g (dex)
[Fe/H] (dex)
v sin i (km s−1)
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
age (Gyr)
Apparent V (mag)
Spec. parallax
Transit model
P (days)
T0 (BJD-2454833)
RP /R(cid:63) (%)
RP (R⊕)
a (AU)
Sinc (S⊕)
Teq (K)
Circular RV model
P (days)
T0 (BJD)
K (m s−1)
γHIRES (m s−1)
γHARPS (m s−1)
γPFS (m s−1)
γFIES (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,PFS (m s−1)
σjit,FIES (m s−1)
MP (M⊕)
ρ (g cm−3)
Eccentric RV model (adopted)
P (days)
T0 (BJD)
K (m s−1)
e
γHIRES (m s−1)
γHARPS (m s−1)
γPFS (m s−1)
γFIES (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
σjit,HARPS (m s−1)
σjit,PFS (m s−1)
σjit,FIES (m s−1)
MP (M⊕)
ρ (g cm−3)
329 ± 10
fixed
fixed 2456940 ± 16
12.7 ± 1.3
17.4 ± 2.8
2.1 ± 2.2
−3.4 ± 3.9
1.3 ± 3.6
2.7 ± 3.1
0 (fixed)
7.42 ± 0.86
6.7 ± 1.4
5.8 ± 1.4
7.2 ± 1.6
37.3 ± 4.3
1.10+0.44−0.31
–
–
327.2 ± 9.8
fixed
fixed 2456614 ± 25
13.8 ± 1.4
18.4 ± 2.9
0.152+0.084
−0.068
1.9 ± 2.2
24486.6 ± 3.9
−1.7 ± 3.5
24574.5 ± 3.0
0 (fixed)
7.57 ± 0.86
6.3 ± 1.4
5.0 ± 1.5
6.7 ± 1.6
39.8 ± 4.4
1.17+0.47−0.32
–
–
Note-A: This work; B: Crossfield et al. (2016). We model a large amplitude (≈ 20m s−1) stellar activity signal by introducing an
additional circular Keplerian.
11
is RP /R(cid:63) = 2.82 ± 0.17%. For an additional star to
significantly alter the observed transit depth requires a
flux ratio, F2/F1 ≈ σ
/ (RP /R(cid:63))2 ≈ 0.10
or ∆V ≈ 2.5 mag (Ciardi et al. 2015).
(RP /R(cid:63))2(cid:17)
(cid:16)
A search for secondary spectral lines in the HIRES
template spectrum (Kolbl et al. 2015) revealed no ad-
ditional stellar companions having ∆V (cid:46) 5 mag and
∆RV (cid:38) 15 km s−1, corresponding to a physical separa-
tion of (cid:46) 1 AU.
We obtained high-resolution speckle imaging at
692 nm with the Differential Speckle Survey Instru-
ment (DSSI; Horch et al. 2012), a visitor instrument
used at the Gemini-North 8.1-m telescope on 2016-01-
13. The observations revealed no additional compan-
ions with ∆V < 2.5 mag in the K2 photometric aper-
ture down to separations of 80 mas, or ≈ 32 AU in
projected separation at the distance of K2-108.4 Ad-
ditional high-resolution imaging with DSSI at 880 nm
and Keck/NIRC2 at K-band also revealed no additional
companions. Of the three high-resolution images, the
DSSI image at 692 nm (shown in Figure 8), provides
the tightest contrast curve in the Kepler bandpass. All
imaging datasets are available on the Exoplanet Follow-
up Observing Program (ExoFOP) website.5
A stellar companion having a ≈ 1–40 AU would
have evaded the aforementioned follow-up observations.
However, such an object would induce a reflex accelera-
tion on K2-108, which would be easily detectable by our
RVs:
γ ≈ GM(cid:63)
a2 ≈ 120 m s−1 yr−1
Over the ∼1 year observation baseline, we see only weak
evidence for a long-term acceleration ( γ = −11.0 ±
2.3 m s−1 yr−1). To be consistent with the observed γ,
a companion at ≈ 40 AU would be (cid:46) 0.1 M(cid:12) and much
too faint to alter the derived planet radius. At smaller
separations, the limits on the mass of a putative com-
panion grow more stringent.
(cid:18) M(cid:63)
(cid:19)(cid:16) a
M(cid:12)
40 AU
(cid:17)−2
.
Figure 4. Searches for Keplerian signals in the radial ve-
locity (RV) time series of K2-39 (see Section 4.3). The top
three panels show searches for Keplerian signatures in the
HIRES RV time series of K2-39 using a Two-Dimensional
Keplerian Lomb-Scargle (2DKLS) periodogram. The bot-
tom panel shows the S-values for K2-39, which traces stellar
activity. First panel: The 2DKLS search of the HIRES data
reveals a peak at P = 4.6 d, which corresponds to K2-39b,
the transiting planet. Second panel: Change in χ2 when
comparing a 2-planet fit to a 1-planet fit (power). There are
several putative signals that are more significant then the
10% eFAP threshold plotted in red (Howard & Fulton 2016).
However, the ≈330 d signal is associated with an activity
cycle, seen in the S-values. We fit this activity signal with a
Keplerian, and perform a final 2DKLS search for additional
planets, shown in the third panel. All other potential signals
fall below the 10% eFAP threshold.
5. DISCUSSION
5.1. Mass and radius
Here, we put the four sub-Saturns presented in this
paper in the context of other planets in their size
class. Starting with the database of exoplanet properties
hosted at the NASA Exoplanet Archive (NEA; Akeson
et al. 2013), we constructed a list of sub-Saturns from
the literature that have densities measured to 50% or
better. We supplemented the list with additional mea-
4 d ≈ 400 pc computed according to the same technique used
for K2-39 (see Section 3).
5 https://exofop.ipac.caltech.edu/k2/
K2-39_hires, 1 planet fit vs. 0 planet fit 4.61.3166.510100Period [Days]0.00.20.40.60.8Power 1 planet fit, RMS=14.71 m s−1, σerr=7.25 m s−1, N=422300240025002600HJD-2455000-60-40-2002040RV [m s−1]Residuals to 1 planet fit, RMS=10.64 m s−12300240025002600HJD-2455000-60-40-200204060RV [m s−1]Phased planet #1, K=13.64 m s−1, e=0.05-0.50.00.5Phase (P=4.608 days, Tc=2570.344)-60-40-2002040RV [m s−1]K2-39_hires, 2 planet fit vs. 1 planet fit 29.9331.019.510100Period [Days]0.00.20.40.60.8Power 2 planet fit, RMS=14.71 m s−1, σerr=7.25 m s−1, N=422300240025002600HJD-2455000-60-40-2002040RV [m s−1]Residuals to 2 planet fit, RMS=7.50 m s−12300240025002600HJD-2455000-40-2002040RV [m s−1]Phased planet #2, K=10.54 m s−1, e=0.05-0.50.00.5Phase (P=29.888 days, Tc=2558.128)-40-2002040RV [m s−1]K2-39_hires, 3 planet fit vs. 2 planet fit (adopted)11.342.223.610100Period [Days]0.00.20.40.60.8Power 2 planet fit, RMS=14.71 m s−1, σerr=7.25 m s−1, N=422300240025002600HJD-2455000-60-40-2002040RV [m s−1]Residuals to 2 planet fit, RMS=7.50 m s−12300240025002600HJD-2455000-40-2002040RV [m s−1]Phased planet #2, K=10.54 m s−1, e=0.05-0.50.00.5Phase (P=29.888 days, Tc=2558.128)-40-2002040RV [m s−1]2300240025002600270028002900BJD - 24548330.140.160.180.200.220.24S-valueS-value Time Series12
Figure 5. Fit to the K2-39 radial velocities (RVs) with with two Keplerians (see Section 4.3). a) Time series of RVs from
HIRES along with FIES, HARPS, and PFS published in Van Eylen et al. (2016b). During the fitting we allowed for an arbitrary
offset between the four instruments to float as a free parameter. The blue line shows the most probable two-Keplerian model.
The long-period Keplerian models out a long period activity cycle, which is also apparent in the S-values. b) Residuals to the
most probable Keplerian model. Panels c) and d) show the phase-folded RVs and the most probable Keplerian model with the
contributions from the other planets removed. The large red circles show the phase-binned RVs. Note that because the posterior
on eb is asymmetric, the eccentricity of the most probable Keperian differs slightly from the value reported in Table 5 (eb =
0.152+0.084
−0.068), which reflects the median of the posterior distribution.
surements that have yet to be ingested into the NEA and
removed a few planets with unreliable measurements.
Including the measurements from this work, we found 19
sub-Saturns that passed our quality cuts and are listed
in Table 7.
In Figure 9, we show the measured masses and sizes
of these planets, highlighting the measurements from
this work. Remarkably, for sub-Saturns, there is lit-
tle correlation between a planet's mass and size. These
planets have a nearly uniform distribution of mass from
MP ≈ 6−60 M⊕. We note that the hottest planets tend
to have higher masses, while cool planets have both high
and low masses. Figure 10 shows planet mass and ra-
dius, cast in terms of planet size and planet density.
The decreasing densities toward larger sizes can be un-
derstood simply in terms of larger radii, given there is no
strong trend of larger planet masses with larger planet
size.
5.2. Planetary Envelope Fraction
One advantage of sub-Saturns is that in this size
range, both heavy elements and low density gaseous en-
velopes contribute significantly to the total planet mass.
A consequence is, to first order, sub-Saturns can be ap-
proximated as two-component planets consisting of a
rocky heavy element core, surrounded by an envelope
of H/He (Lopez & Fortney 2014; Petigura et al. 2016).
Here we use the results of Lopez & Fortney (2014),
which simulated the internal structure and thermal evo-
−40−30−20−1001020304050RV [m s-1]a)FIESHARPSHIRESPFS2015.62015.82016.02016.22016.42016.67300740075007600BJDTDB - 2450000−25025Residualsb)−0.4−0.20.00.20.4Phase−30−20−100102030RV [m s-1]c)Pb = 4.60 daysKb = 14.06 m s-1eb = 0.19 −0.4−0.20.00.20.4Phase−30−20−100102030RV [m s-1]d)Pc = 330.86 daysKc = 19.40 m s-1ec = 0.00 13
Figure 6. Top: K2 light curve of K2-108 showing with the transits of K2-108b labeled with red ticks. Bottom: photometry
phase-folded on the transit ephemeris.
Figure 7. Single Keplerian model to the K2-108 radial velocities (RVs), allowing for eccentricity (see Section 4.4). a) Time
series of RVs from HIRES. The blue line shows the most probable Keplerian model. b) Residuals to the most probable Keplerian
model. c) The phase-folded RVs and the most probable Keplerian model.
.
−40−30−20−1001020304050RV [m s-1]a)HIRES2016.02016.22016.42016.62016.87400745075007550760076507700BJDTDB - 2450000−25025Residualsb)−0.4−0.20.00.20.4Phase−40−30−20−10010203040RV [m s-1]c)Pb = 4.73 daysKb = 21.44 m s-1eb = 0.18 14
Table 6. System parameters of K2-108
Value
Stellar parameters
Identifier
Teff (K)
log g (dex)
[Fe/H] (dex)
v sin i (km s−1)
M(cid:63) (M(cid:12))
R(cid:63) (R(cid:12))
age (Gyr)
Apparent V (mag)
Ref.
EPIC-211736671
5474 ± 60 A
3.99 ± 0.05 A
0.33 ± 0.04 A
< 2 A
−0.053 A
1.121+0.065
1.75 ± 0.14 A
7.8 ± 1.5 A
12.33 ± 0.01 A
Figure 8. Contrast curve of K2-108 (EPIC-211736671) taken
with a narrow-band filter centered at 692 nm using the DSSI
camera (Horch et al. 2012) on the Gemini-N 8m telescope.
The inset image shows the 2×2 arcsec region centered on K2-
108 reconstructed from the speckle image sequence. Putative
companions with contrasts of <2.5 mag which could alter the
inferred planet size are ruled out for separations > 80 mas.
Additional DSSI observations at 880 nm and Keck/NIRC2
observations at K-band also reveal no additional companions
to K2-108. See Section 4.4 for further details.
lution of planets with solar-composition H/He envelopes
atop Earth-composition cores. These simulations were
run over a wide swath of parameter space for planets
with different masses, MP , different envelope fractions,
fenv = Menv/MP , and on orbits receiving different levels
of incident stellar irradiation, Sinc. The model planets
were allowed to evolve over time and Lopez & Fortney
(2014) noted the planet radii at specified intervals. The
result of these simulations is a four-dimensional grid of
planet radius, RP , sampled at various combinations of
(MP , fenv, Sinc, age).
We used this grid to solve for the values of planet
fenv that are consistent with the observed (MP , RP ,
Sinc, and age). We drew MP , RP , and Sinc measured
posterior distributions and interpolated the model grid
in order to derive fenv. We assumed a uniform age of
5 Gyr. With the exception of a few stars analyzed with
asteroseismology, most of the system ages are derived
from isochrone-fitting and are thus uncertain at the ≈2–
3 Gyr level. Fortunately, the derived fenv fraction is not
sensitive to the adopted system age. Adopting a uniform
age of 2 Gyr resulted in typical change in the derived
fenv of ≈ 2%. The resulting values of fenv are listed in
Table 7.
We have made several approximations when comput-
ing fenv. We have ignored the possibility of water or
other volatile ices contributing significantly to heavy el-
ement cores of these planets. Lopez & Fortney (2014),
planet b
4.73401 ± 0.00024 A
2312.0965 ± 0.0019 A
5.28 ± 0.54 A
0.0573 ± 0.0010 A
762 ± 100 A
1446 ± 48 A
Transit model
P (days)
T0 (BJD-2454833)
RP (R⊕)
a (AU)
Sinc (S⊕)
Teq (K)
Circular RV model
K (m s−1)
γHIRES (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
MP (M⊕)
ρ (g cm−3)
Eccentric RV model (adopted)
K (m s−1)
e
γHIRES (m s−1)
γ (m s−1 yr−1)
σjit,HIRES (m s−1)
MP (M⊕)
ρ (g cm−3)
19.4 ± 2.3 A
−4.0 ± 1.6 A
0 (fixed) A
6.0+1.5−1.1 A
55.1 ± 6.8 A
2.05+0.75−0.54 A
21.3 ± 1.4 A
0.180 ± 0.042 A
−2.63 ± 0.91 A
−11.0 ± 2.3 A
2.86+1.01−0.78 A
59.4 ± 4.4 A
2.22+0.77−0.55 A
Note-A: This work
however, showed that including ices in the core, does
not significantly alter the radius-composition relation-
ship for planets in this size range. Because the H/He
envelope represents most of the planet volume, our in-
ferred fenv does not depend sensitively on the detailed
composition of the planet core. We have also assumed
that all the heavy elements are concentrated in the plan-
ets' cores rather than being distributed throughout the
envelope.
5.3. Role of Photo-Evaporation among Sub-Saturns
If photo-evaporation plays a dominant role in sculpt-
ing the gaseous envelopes of sub-Saturns, one might ex-
pect the envelope fraction, fenv, to correlate with the
energy it receives from its host star. Figure 11 shows the
total planet mass (MP ), and envelope fraction (fenv) as a
function of the blackbody equilibrium temperature. As
a matter of convenience, we include core mass (Mcore)
and envelope mass (Menv) which can be trivially com-
puted from MP and fenv. We do not observe a strong
one-to-one correlation envelope fraction and equilibrium
temperature. We do, however, observe that cool planets
span a large range of fenv ≈ 10 − 50% while hot planets
span a more narrow range of fenv ≈ 10 − 20%. Perhaps
photo-evaporation excludes planets from occupying cer-
tain domains in the fenv–Teq plane.
To further explore the possible role of photo-
evaporation, we considered two quantities that are more
directly related to a planets susceptibility to photo-
evaporation: XUV heating and planet binding energy.
XUV heating is the total lifetime-integrated XUV flux
incident at a planet's orbit multiplied by the planet's
current cross section. Figure 12, which is updated from
(Lopez & Fortney 2014), shows the planet XUV heating
(in ergs) vs. planet binding energy (in ergs). This sort
of diagram has been used as supporting evidence for the
role of photo-evaporation in sculpting the envelopes of
highly-irradiated sub-Neptunes (Lecavelier Des Etangs
2007; Lopez et al. 2012; Owen & Wu 2013). The dashed
line shows the envelope survival threshold predicted
by photo-evaporation and thermal evolution models in
Lopez et al. (2012), and the absence of planets with
gaseous envelopes above this line suggests that planets
near this threshold have experienced photo-evaporation.
While a few sub-Saturns lie close to this threshold, most
lie well below it, indicating that they are sufficiently
massive or are on wide enough orbits to be immune to
significant photo-evaporation.
5.4. Stellar Metallicity and Planet Metallicity
One may interpret present day stellar metallicity as
a proxy for the metal-enrichment of the protoplanetary
disk because the disk and star formed from the same
molecular cloud. However, the mean metallicity of the
protoplanetary disk may be different from the local disk
metallicity at the location of planet formation. With
this caveat in mind, we nonetheless looked for corre-
lations between host star metallicity and the observed
properties of sub-Saturns.
In contrast to equilibrium temperatures we observe a
stronger set of correlations between the planetary prop-
erties of sub-Saturns and host star metallicity, [Fe/H].
Figure 13 shows MP , fenv, Mcore, and Menv against
[Fe/H]. Sub-Saturns orbiting metal-rich stars tend to
be more massive.
Interestingly, both the planetary
core mass, Mcore, and envelope mass, Menv, appear to
increase with stellar metallicity. This is understand-
able, however, given that planets with more massive
cores should generally be able to accrete larger gaseous
envelopes before their disks dissipate (Lee & Chiang
2015b). This suggests that the metal-rich hosts had
more solids available in their disk allowing those sub-
Saturns to form more massive cores; and that these
larger cores were then able accrete more massive en-
velopes.
15
It is worthwhile to compare the observed MP –[Fe/H]
trend to the stellar metallicity distribution of Kepler
planet hosts. Buchhave et al. (2012) observed that plan-
ets smaller than ≈ 4 R⊕ are found around stars of wide-
ranging metallicities (−0.5 (cid:46) [Fe/H] (cid:46) +0.5 dex) while
larger planets are typically found around more metal-
rich stars (−0.2 (cid:46) [Fe/H] (cid:46) +0.5 dex). This is consis-
tent with our sample, constructed from planets found
by Kepler, K2 , and other surveys. Stellar metallicity
of > −0.2 dex seems to be an important criterion for
forming sub-Saturns. However, the additional metals in
the disk seem to result in more massive final planets.
The path by which increased metallicity produces
more massive sub-Saturns could proceed in one of two
ways: (1) disks with more solid material form substan-
tially more massive planet cores which grow smoothly
into more massive planets or (2) disks with more solid
material form slightly more massive cores which perturb
neighboring planets causing collisions and mergers. We
have a slight preference for the latter explanation, given
the eccentricity distribution of sub-Saturns, explored in
Section 5.5. It is also plausible that metal-enriched disks
could form planet cores more quickly, allowing for a long
period of gas accretion. However, we see no evidence of
this given the absence of a strong correlation between
fenv and host star metallicity.
In addition to the trends with stellar metallicity shown
in Figure 13, it is also interesting to examine whether
there are trends in the heavy element abundances of
sub-Saturns after controlling for the dependence on stel-
lar metallicity. Recently, Thorngren et al. (2016) ex-
amined planet metal mass fraction, ZP , for 47 planets
having MP ≈ 30 − 3000 M⊕ relative to the heavy el-
ement fraction of their parent stars, Z(cid:63) = Z(cid:12)10[Fe/H].
ZP was computed via ZP = Mcore/Menv using thermal
evolution models similar to those used here. Thorngren
et al. (2016) observed an anti-correlation between MP
and ZP /Z(cid:63). They argued that such an anti-correlation
can be understood in terms of traditional core-accretion
formation theory if one assumes that planets below the
isolation mass are able accrete all of the solids in their
isolation zone, typically ≈3.5 Hill-radii (Lissauer 1993),
but not all their gas. Given these assumptions, Equa-
tion 9 of Thorngren et al. (2016) predicts the planetary
metal enrichment:
(cid:18) MP
(cid:19)−2/3
ZP
Z(cid:63)
Q−1
H
a
.
M(cid:63)
= 3fH fe
(1)
Here, fH ∼ 3.5 is the approximate number of Hill-radii
from which a planet can effectively accrete solids, fe ∼ 1
is an enrichment factor to allow for metal enhancement
due to radial drift by solids, H is the disk scale height,
a is the semi-major axis, and Q is Toomre disk instabil-
ity parameter (Toomre 1964). Thorngren et al. (2016)
16
found that Equation 1 can reproduce the observed trend
between MP and ZP /Z(cid:63) if fH = 3.5, fe = 1, and Q = 5.
Figure 14 compares our sample of sub-Saturns to the
predictions of Equation 1. Sub-Saturns are a valuable
laboratory for testing the physics of envelope accretion
because they are larger than the sub-Neptunes, which
typically have fenv (cid:46) 10%, and the gas giants, which
are nearly entirely envelope (fenv ∼ 100%). Follow-
ing Thorngren et al. (2016), we approximate the plane-
tary metal abundance by setting ZP = Mcore/MP . For
the most massive Sub-Saturns (having MP /M(cid:63) (cid:38) 10−4)
where our sample overlaps with the Thorngren et al.
(2016) sample, we find good agreement with the predic-
tions of Equation 1, as shown by the dotted line. Below
MP /M(cid:63) ∼ 10−4, however, we find a significant increase
in dispersion below this relation, with many planets be-
ing significantly less enriched in metals than expected
from Equation 1.
One interpretation is that this increase in scatter sim-
ply reflects the natural transition between giant plan-
ets, which essentially accrete all of the heavy elements
in their feeding zones, and low mass planets which do
not. Equation 1, assumes that the solids in the disk
have fully decoupled from the gas, and that a planet
can successfully accrete all of the solids near its isola-
tion zone. At lower cores masses, gravitational focus-
ing becomes more important, gas damping of planetesi-
mals eccentricities becomes more efficient, and collision
and growth timescales become longer. All of these fac-
tors mean less massive cores may not accrete all solids
with in ≈ 3.5 Hill-radii, which may instead be dispersed
or incorporated into other planets in compact, multi-
planet systems. In summary, we examined whether the
correlation between planet metal-enrichment and planet
masses observed by Thorngren et al. (2016) for planets
having MP = 30–3000 M⊕ is present among the sub-
Saturns. We do not observe a strong correlation, indi-
cating a possible transition in the formation pathways
of planets with MP (cid:46) 30 M⊕.
5.5. Eccentricity and Planet Multiplicity
The fits to K2-27, K2-39, and K2-108 RVs favored non-
zero orbital eccentricities of 0.251 ± 0.088, 0.152+0.084
−0.068,
and 0.180±0.042, respectively. Given that these planets
are on short orbital periods of 4.6–6.8 d, respectively, it
is worthwhile to consider the extent to which tides are
expected to damp away eccentricity. The timescale for
eccentricity damping (Goldreich & Soter 1966) is given
by
(cid:18) Q(cid:48)
(cid:19)(cid:18) MP
(cid:19)(cid:18) a
τe =
Here, n =(cid:112)GM(cid:63)/a3 is the mean motion, and Q(cid:48), the
modified tidal quality factor, is given by Q(cid:48) = 3Q/2k2,
(2)
M(cid:63)
RP
n
.
(cid:19)5
4
63
where Q is the specific dissipation function and k2 is the
Love number (see Goldreich & Soter 1966; Murray &
Dermott 1999; Mardling & Lin 2004). Q(cid:48) is quite uncer-
tain even for planets in the Solar System. As a point of
reference Lainey et al. (2015) give Q(cid:48) ≈ 6, 000 − 18, 000
for Saturn, based on Cassini ranging data. Tittemore
& Wisdom (1989, 1990) give Q ≈ 11, 000 − 39, 000
for Uranus based on studies of the Uranian satilites,
which translates to Q(cid:48) ≈ 165, 000 − 585, 000, adopt-
ing k2 = 0.104 from (Gavrilov & Zharkov 1977). Q(cid:48)
is even more uncertain for sub-Saturns which have no
Solar System analogs. Here, we adopt Q(cid:48) = 105 for the
sub-Saturns, with the understanding that this estimate
is only good to order of magnitude. Re-writing Equa-
tion 2,
(cid:17)−1
(cid:1)(cid:16) MP
(cid:17)(cid:16) M(cid:63)
(cid:17)−5
(cid:1)5(cid:16) RP
10 M⊕
M(cid:12)
.
4R⊕
a
0.05 AU
τe ∼ 1.8 Gyr
(cid:17)(cid:0) P
10 d
(cid:16) Q(cid:48)
×(cid:0)
105
For K2-27b, we find τe ∼ 10 Gyr, comparable to the
age of the system, suggesting that if this eccentricity was
caused by planet-planet scattering early in the star's life-
time, the eccentricity could persist to the present day.
K2-39b and K2-108b are slightly larger than K2-27b and
also orbit closer to their host stars. Because the circular-
ization timescale is a strong function of a/RP , they have
substantially shorter τe of ∼ 0.6 Gyr and ∼ 2 Gyr, re-
spectively. These circularization timescales are formally
shorter than the age of their host stars and present some
challenges for understanding any present day eccentric-
ities. This tension could be resolved if Q(cid:48) is ∼ 106 as
opposed to the assumed value of ∼ 105. Eccentric or-
bits could also be maintained by additional, yet unde-
tected planets. Deming et al. (2007) proposed such an
explanation for GJ436b, another short-period eccentric
sub-Saturn (see Table 7). Assuming Q(cid:48) ∼ 105, τe for
GJ436b is only ∼ 0.1 Gyr. While no additional planets
have been detected in the GJ436 system, Batygin et al.
(2009) showed this explanation to be plausible in the
context of secular theory.
In Table 7, we also included the measured eccentric-
ities, when available. We broke the sample into low (e
< 0.1), moderate (e > 0.1), and poorly-constrained ec-
centricities. In order for a planet to be included in the
low/moderate eccentricity bins, its entire 1σ eccentric-
ity confidence interval must be below/above 0.1. Planets
with eccentricity upper limits or constraints straddling
0.1 are fall in the poorly-constrained category. These
different eccentricity categories are color-coded in Fig-
ures 11, 13, and 15.
We observed that the highest-mass planets were of-
ten the only detected planet in the system. Figure 15
shows MP , fenv, Mcore, and Menv as a function of the
total number of detected planets in the system. There
is a steady decline in the mass of sub-Saturns as over-
all multiplicity increases. The planets with moderate
eccentricities are confined to apparently single systems.
These high-mass singles could have originally had neigh-
bors, but were in a dynamically unstable architecture.
Such instabilities would eventually lead to close en-
counters and planet-planet scattering. Because these
planets are so deep in the potential wells of their host
stars, these scattering events would likely lead to merg-
ers as opposed to ejections from the system. The maxi-
mum velocity a planet can impart to its neighbor is the
escape velocity,
and if vesc is smaller than the orbital velocity,
vesc = 11.2 km s−1
vorb = 30 km s−1
(cid:18) MP
(cid:19)1/2(cid:18) RP
(cid:19)1/2(cid:16) a
(cid:18) M(cid:63)
M⊕
R⊕
M(cid:12)
1 AU
(cid:19)−1/2
(cid:17)−1/2
.
,
single scattering events cannot lead to ejections. For the
sub-Saturns in Table 7, vesc/vorb ≈ 0.1− 0.3.6 Any pre-
vious dynamical instabilities would likely lead to planet
mergers, increasing their total mass. The present-day
17
eccentricities of the more massive sub-Saturns may be a
relic of previous scattering and merging events.
It is worth considering the biases associated with the
different techniques by which planet mass and eccen-
tricity are measured and whether they could be re-
sponsible for the observed mass-multiplicity-eccentricity
trends. TTV measurements require multi-planet system
and thus do not contribute to any points in the NP = 1
bin in Figure 15. Constraining e < 0.1 is challenging
with RVs given that one is trying to measure slight devi-
ations from sinusoidal RV curves. Therefore, limitations
of the RV and TTV techniques might explain why there
are no single planets with secure eccentricity measure-
ments of < 0.1.
However, these observational biases cannot explain
why planets in multi-planet systems are low mass and
preferentially circular. Previous studies (e.g. Weiss &
Marcy 2014) have noted that planets with TTV mass
measurements are typically less massive than planets
with RV mass measurements. TTVs, however, are not
blind to high-mass planets; such planets would produce
larger TTVs. The lack of high-mass, high-eccentricity
planets in multi-planet systems is likely the result of
dynamical instabilities. Planets in such systems would
likely perturb one another and merge, resulting high-
mass planets in low-multiplicity systems.
Name
NP
Kepler-4 b
GJ 436 b
Kepler-11 e
Kepler-413 b
K2-27 b
Kepler-223 e
HAT-P-11 b
K2-32 b
Kepler-25 c
Kepler-223 d
K2-108 b
Kepler-18 c
K2-24 b
K2-39 b
Kepler-101 b
Kepler-87 c
HATS-7 b
HAT-P-26 b
1
1
6
1
1
4
1
3
3
4
1
3
2
1
2
2
1
1
Table 7. Sub-Saturns with well-measured densities
RP
R⊕
4.00+0.21−0.21
4.17+0.17−0.17
4.19+0.07−0.09
4.35+0.10−0.10
4.48+0.23−0.23
4.60+0.27−0.41
4.73+0.16−0.16
5.13+0.28−0.28
5.20+0.09−0.09
5.24+0.26−0.45
5.28+0.54−0.54
5.49+0.26−0.26
5.68+0.56−0.56
5.71+0.63−0.63
5.77+0.85−0.79
6.14+0.29−0.29
6.31+0.52−0.38
6.33+0.81−0.36
MP
M⊕
24.5+3.8−3.8
22.1+2.3−2.3
8.0+1.5−2.1
51.0+22.0−21.0
30.9+4.6−4.6
4.8+1.4−1.2
25.7+2.9−2.9
16.5+2.7−2.7
24.6+5.7−5.7
8.0+1.5−1.3
59.4+4.4−4.4
17.3+1.9−1.9
21.0+5.4−5.4
39.7+4.6−4.6
51.1+5.1−4.7
6.4+0.8−0.8
38.1+3.8−3.8
18.8+2.2−2.2
ρ
g cm−3
2.09+0.53−0.44
1.67+0.30−0.26
0.60+0.15−0.14
2.40+1.00−1.00
1.88+0.46−0.38
0.27+0.11−0.09
1.34+0.22−0.20
0.67+0.18−0.15
0.96+0.24−0.23
0.30+0.10−0.07
2.21+0.90−0.59
0.57+0.12−0.10
0.62+0.31−0.22
1.16+0.54−0.34
1.46+0.90−0.50
0.15+0.03−0.03
0.83+0.23−0.17
0.40+0.15−0.10
e
· · ·
0.1383+0.0002
−0.0002
0.0120+0.0060
−0.0060
0.1185+0.0018
−0.0017
0.2510+0.0880
−0.0880
0.0510+0.0190
−0.0190
0.1980+0.0460
−0.0460
· · ·
· · ·
0.0370+0.0180
−0.0170
0.1800+0.0420
−0.0420
· · ·
· · ·
0.1500+0.0760
−0.0760
0.0860+0.0800
−0.0590
0.0390+0.0120
−0.0120
· · ·
0.1240+0.0600
−0.0600
Teq
K
1597
659
630
348
910
944
861
815
1018
1040
1440
979
766
1689
1547
440
1070
981
[Fe/H]
fenv
%
dex
6.7+1.3−1.4 +0.17
· · ·
12.6+1.9−1.9
14.9+0.7−0.7 −0.04
· · ·
10.8+3.8−2.0
13.9+2.4−2.5 +0.13
16.5+2.5−2.6 +0.06
17.1+1.5−1.5 +0.31
22.5+3.0−3.1 −0.02
21.4+1.3−1.3 −0.04
22.3+3.4−3.2 +0.06
16.0+4.8−5.4 +0.33
25.7+2.6−2.8 +0.19
28.4+7.4−6.2 +0.42
18.1+5.2−4.8 +0.43
19.5+7.9−6.9 +0.33
35.6+3.4−3.7 −0.17
33.3+6.1−5.8 +0.25
35.8+6.9−7.5 −0.04
6 We have excluded Kepler-413 b, because it is a circumbinary
planet with more complex criteria for ejection.
Table 7 continued
18
Table 7 (continued)
Name
NP
CoRoT-8 b
Kepler-56 b
Kepler-18 d
Kepler-79 d
K2-24 c
1
3
3
4
2
RP
R⊕
6.39+0.22−0.22
6.51+0.29−0.28
6.98+0.33−0.33
7.16+0.13−0.16
7.82+0.72−0.72
MP
M⊕
69.9+9.5−9.5
22.1+3.9−3.6
16.4+1.4−1.4
6.0+2.1−1.6
27.0+6.9−6.9
ρ
g cm−3
1.47+0.27−0.25
0.44+0.10−0.09
0.26+0.05−0.04
0.09+0.03−0.03
0.31+0.14−0.10
e
· · ·
· · ·
· · ·
0.0250+0.0590
−0.0230
· · ·
Teq
K
844
1479
784
626
605
[Fe/H]
fenv
%
dex
32.2+3.5−3.1 +0.30
26.3+2.4−2.4 +0.37
44.5+3.2−3.6 +0.19
42.6+2.5−3.3 −0.02
57.3+9.1−9.9 +0.42
Note-List of sub-Saturns having densities measured to 50% or better, assembled from the NASA Exoplanet
Archive, this work, and other sources. NP -Total number of detected planets in the system, e-orbital
eccentricity, we do not report eccentricity if only upper limits are available. Teq refers to the blackbody
temperature (i.e. assuming zero albedo). fenv-"envelope fraction" fraction of planet's mass in H/He in the
two-component modeling of planet mass and radius, described in Section 5.2. Notes on individual systems:
Kepler-4 b-Borucki et al. (2010a); GJ 436 b-Butler et al. (2004); Kepler-11 e-Lissauer et al. (2013);
Kepler-413 b-Kostov et al. (2014); K2-27 b-This work; Kepler-223 e-Mills et al. (2016); HAT-P-11
b-Bakos et al. (2010); K2-32 b-This work; Kepler-25 c-Marcy et al. (2014); Kepler-223 d-Mills et al.
(2016); K2-108b-This work; Kepler-18 c-Cochran et al. (2011); K2-24 b-Petigura et al. (2016); K2-39
b-This work; Kepler-101 b-Bonomo et al. (2014); Kepler-87 c-Ofir et al. (2014); HATS-7 b-Bakos et al.
(2015); HAT-P-26 b-Hartman et al. (2011); CoRoT-8 b-Bordé et al. (2010), adopted 3± 1 Gyr; Kepler-56
b-Huber et al. (2013); Kepler-18 d-Cochran et al. (2011); Kepler-79 d-Jontof-Hutter et al. (2014); K2-24
c-Petigura et al. (2016).
6. CONCLUSIONS
We presented radial velocity measurements of four sys-
tems hosting sub-Saturn planets observed by the K2
mission: K2-27, K2-32, K2-39, and K2-108. These RVs
enabled mass measurements of 16% or better and de-
tailed analysis of the planetary heavy element fraction.
Despite the similar sizes of the planets, their masses
range from 16–60 M⊕, implying widely different core
and envelope masses. Despite the differences in the
masses of these planets, the fraction of their mass in
H/He is similar ≈80%. This trend is seen in the popu-
lation of ≈20 sub-Saturns with well-measured masses.
Sub-Saturns as a class of planets show a remarkable
diversity in mass MP = 6–60 M⊕, with little dependence
on planet size. We observe a strong correlation between
stellar metallicity and planet mass. This implies that
metal-rich disks produce more, or more massive, cores.
Finally, we observe a tendency of the most massive sub-
Saturns to have moderate eccentricities and to reside
in apparently single systems. Future observational and
theoretical work will further illuminate these mysterious
planets, absent in our own Solar System.
We thank Konstantin Batygin and John Livingston
for helpful discussions. E. A. P. acknowledges support
from a Hubble Fellowship grant HST-HF2-51365.001-A
awarded by the Space Telescope Science Institute, which
is operated by the Association of Universities for Re-
search in Astronomy, Inc.
for NASA under contract
NAS 5-26555. This work has made use of data from
the European Space Agency (ESA) mission Gaia, pro-
cessed by Gaia Data Processing and Analysis Consor-
tium (DPAC). Funding for the DPAC has been pro-
vided by national institutions, in particular the insti-
tutions participating in the Gaia Multilateral Agree-
ment. Some of the data presented herein were obtained
at the W. M. Keck Observatory (which is operated as a
scientific partnership among Caltech, UC, and NASA).
We thank the Caltech and NASA Keck Time Alloca-
tion Committees for providing HIRES time. This work
included observations obtained at the Gemini Observa-
tory, which is operated by the Association of Universi-
ties for Research in Astronomy, Inc., under a coopera-
tive agreement with the NSF on behalf of the Gemini
partnership: the National Science Foundation (United
States), the National Research Council (Canada), CON-
ICYT (Chile), Ministerio de Ciencia, Tecnología e Inno-
vación Productiva (Argentina), and Ministério da Ciên-
cia, Tecnologia e Inovação (Brazil).The authors wish to
recognize and acknowledge the very significant cultural
role and reverence that the summit of Maunakea has
always had within the indigenous Hawaiian community.
We are most fortunate to have the opportunity to con-
duct observations from this mountain.
Software: Numpy/Scipy (Van Der Walt et al.
2011), Matplotlib (Hunter 2007), Pandas
(McK-
inney
2010),
(Astropy Collaboration
et al. 2013),
(Goodman & Weare 2010;
Foreman-Mackey et al. 2013), SME (Brewer et al.
(Aigrain
2015),
et al. 2015),
radvel
(https://github.com/California-Planet-Search/radvel),
batman (Kreidberg 2015),
Astropy
emcee
isochrones
(Morton 2015),
k2sc
19
Figure 9. Top: Masses and radii of sub-Saturns having densities measured to 50% or better. Planets of different size classes
with comparable density precision are shown as gray points for context. For the sub-Satruns, symbol colors correspond to the
blackbody equilibrium temperature. The symbol shapes correspond to the method by which planet masses were measured:
radial velocities (RVs), transit-timing variations (TTVs), or a combined analysis (RVs+TTVs). For sub-Saturns, we note almost
no correlation between planet mass and planet size. Bottom: a zoomed in view of the top panel, focusing on sub-Saturns. See
Section 5.1 for additional details.
13103010030010003000Planet Mass (Earth-masses)1234681020Planet Size (Earth-radii)RVTTVRV+TTV20040060080010001200140016001800Equilibrium Temp (K)31030100Planet Mass (Earth-masses)456789Planet Size (Earth-radii)RVTTVRV+TTV20040060080010001200140016001800Equilibrium Temp (K)20
Figure 10. Same as Figure 9 but showing mean planet density as a function of planet size. For planets of a given size, there is
a diversity of densities due to the diversity in planet mass. While the hottest planets seem to have high densities for their size,
cool planets span a wide range of density.
45678Planet Size (Earth-radii)0.050.070.10.20.30.40.50.7123Planet Density (g/cc)RVTTVRV+TTV20040060080010001200140016001800Equilibrium Temp (K)21
Figure 11. Panels a–d show the planet mass (MP ), envelope fraction (Menv/MP ), core mass (Mcore), and envelope mass (Menv)
as a function of planet blackbody equilibrium temperature (Teq). We observe an absence of planets having high envelope
fractions (fenv (cid:38) 0.25) at high equilibrium temperatures as expected from photo-evaporation. However, the lack of a strong
trend between Teq and fenv suggests that photo-evaporation has not significantly sculpted the majority of sub-Saturns shown.
Planets where eccentricity has been constrained to be less than or greater than 0.1 are colored blue and red, respectively.
k2phot (https://github.com/petigura/k2phot)
REFERENCES
Aigrain, S., Hodgkin, S. T., Irwin, M. J., Lewis, J. R., &
Brewer, J. M., Fischer, D. A., Basu, S., Valenti, J. A., &
Roberts, S. J. 2015, MNRAS, 447, 2880
Akeson, R. L., Chen, X., Ciardi, D., et al. 2013, PASP, 125, 989
Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al.
2013, A&A, 558, A33
Bakos, G. Á., Torres, G., Pál, A., et al. 2010, ApJ, 710, 1724
Bakos, G. Á., Penev, K., Bayliss, D., et al. 2015, ApJ, 813, 111
Batygin, K., Laughlin, G., Meschiari, S., et al. 2009, ApJ, 699, 23
Bodenheimer, P., Hubickyj, O., & Lissauer, J. J. 2000, Icarus,
143, 2
Bonomo, A. S., Sozzetti, A., Lovis, C., et al. 2014, ArXiv
e-prints, arXiv:1409.4592
Bordé, P., Bouchy, F., Deleuil, M., et al. 2010, A&A, 520, A66
Borucki, W. J., Koch, D. G., Brown, T. M., et al. 2010a, ApJL,
713, L126
Borucki, W. J., Koch, D., Basri, G., et al. 2010b, Science, 327,
977
Piskunov, N. 2015, ApJ, 805, 126
Brewer, J. M., Fischer, D. A., Valenti, J. A., & Piskunov, N.
2016, ApJS, 225, 32
Buchhave, L. A., Latham, D. W., Johansen, A., et al. 2012,
Nature, 486, 375
Butler, R. P., Vogt, S. S., Marcy, G. W., et al. 2004, ApJ, 617,
580
Ciardi, D. R., Beichman, C. A., Horch, E. P., & Howell, S. B.
2015, ApJ, 805, 16
Cochran, W. D., Fabrycky, D. C., Torres, G., et al. 2011, ApJS,
197, 7
Crossfield, I. J. M., Ciardi, D. R., Petigura, E. A., et al. 2016,
ApJS, 226, 7
Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP,
120, 531
Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, VizieR
Online Data Catalog, 2246
25050075010001250150017502000Equilibrium Temp (K)31030100MP (Earth-masses)25050075010001250150017502000Equilibrium Temp (K)0.00.20.40.60.81.0Menv /MPe > 0.1e < 0.1e uncert.25050075010001250150017502000Equilibrium Temp (K)0.3131030100Mcore (Earth-masses)25050075010001250150017502000Equilibrium Temp (K)0.3131030100Menv (Earth-masses)22
Figure 12. The total lifetime XUV heating planets receive at their orbit vs. their current binding energy. Sub-Saturns are
color-coded by their current gaseous envelope fraction, while all other planets <100 M⊕ are shown in grey. The dashed line
meanwhile shows the predicted photo-evaporation threshold from Lopez et al. (2012). This indicates that only a few of the
sub-Saturns in this sample have likely been strongly affected by photo-evaporation.
Dai, F., Winn, J. N., Albrecht, S., et al. 2016, ApJ, 823, 115
Deming, D., Harrington, J., Laughlin, G., et al. 2007, ApJL, 667,
L199
Jontof-Hutter, D., Lissauer, J. J., Rowe, J. F., & Fabrycky, D. C.
2014, ApJ, 785, 15
Kolbl, R., Marcy, G. W., Isaacson, H., & Howard, A. W. 2015,
Dotter, A., Chaboyer, B., Jevremović, D., et al. 2008, ApJS, 178,
AJ, 149, 18
89
Eastman, J., Gaudi, B. S., & Agol, E. 2013, PASP, 125, 83
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J.
2013, PASP, 125, 306
Kostov, V. B., McCullough, P. R., Carter, J. A., et al. 2014,
ApJ, 784, 14
Kreidberg, L. 2015, ArXiv e-prints, arXiv:1507.08285
Lainey, V., Jacobson, R. A., Tajeddine, R., et al. 2015, ArXiv
Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766,
e-prints, arXiv:1510.05870
81
Gavrilov, S. V., & Zharkov, V. N. 1977, Icarus, 32, 443
Goldreich, P., & Soter, S. 1966, Icarus, 5, 375
Goodman, J., & Weare, J. 2010, Communications in Applied
Mathematics and Computational Science, 5, 65
Lecavelier Des Etangs, A. 2007, A&A, 461, 1185
Lee, E. J., & Chiang, E. 2015a, ArXiv e-prints, arXiv:1510.08855
-. 2015b, ApJ, 811, 41
Lissauer, J. J. 1993, ARA&A, 31, 129
Lissauer, J. J., Marcy, G. W., Rowe, J. F., et al. 2012, ApJ, 750,
Hartman, J. D., Bakos, G. Á., Kipping, D. M., et al. 2011, ApJ,
112
728, 138
Lissauer, J. J., Jontof-Hutter, D., Rowe, J. F., et al. 2013, ApJ,
Horch, E. P., Howell, S. B., Everett, M. E., & Ciardi, D. R. 2012,
770, 131
AJ, 144, 165
Howard, A. W., & Fulton, B. J. 2016, PASP, 128, 114401
Howard, A. W., Johnson, J. A., Marcy, G. W., et al. 2010, ApJ,
721, 1467
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS,
201, 15
Howard, A. W., Marcy, G. W., Fischer, D. A., et al. 2014, ApJ,
794, 51
Howell, S. B., Sobeck, C., Haas, M., et al. 2014, PASP, 126, 398
Huber, D., Carter, J. A., Barbieri, M., et al. 2013, Science, 342,
331
Huber, D., Bryson, S. T., Haas, M. R., et al. 2016, ApJS, 224, 2
Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2005, Icarus,
179, 415
Hunter, J. D. 2007, Computing in Science and Engineering, 9, 90
Isaacson, H., & Fischer, D. 2010, ApJ, 725, 875
Lopez, E. D., & Fortney, J. J. 2014, ApJ, 792, 1
Lopez, E. D., Fortney, J. J., & Miller, N. 2012, ApJ, 761, 59
Mamajek, E. E. 2009, in American Institute of Physics
Conference Series, Vol. 1158, American Institute of Physics
Conference Series, ed. T. Usuda, M. Tamura, & M. Ishii, 3–10
Marcy, G. W., & Butler, R. P. 1992, PASP, 104, 270
Marcy, G. W., Isaacson, H., Howard, A. W., et al. 2014, ApJS,
210, 20
Mardling, R. A., & Lin, D. N. C. 2004, ApJ, 614, 955
McKinney, W. 2010, in Proceedings of the 9th Python in Science
Conference, ed. S. van der Walt & J. Millman, 51 – 56
Mills, S. M., Fabrycky, D. C., Migaszewski, C., et al. 2016,
Nature, 533, 509
Mordasini, C., Alibert, Y., Benz, W., & Naef, D. 2008, 398, 235
Morton, T. D. 2015, isochrones: Stellar model grid package,
Astrophysics Source Code Library, , , ascl:1503.010
23
Figure 13. Same as Figure 15 but showing planet mass (MP ), envelope fraction (Menv/MP ), core mass (Mcore), and envelope
mass (Menv) as a function of stellar metallicity, [Fe/H]. We observe a correlation between metal-rich host stars and more massive
sub-Saturns. The host star metallicity does not correlate with fenv, suggesting that disk metallicity is not the only factor that
affects the final add-mixture of envelope and solids that comprise these planets.
Murray, C. D., & Dermott, S. F. 1999, Solar system dynamics
Ofir, A., Dreizler, S., Zechmeister, M., & Husser, T.-O. 2014,
A&A, 561, A103
701, 1732
O'Toole, S. J., Jones, H. R. A., Tinney, C. G., et al. 2009, ApJ,
Owen, J. E., & Wu, Y. 2013, ApJ, 775, 105
Petigura, E. A., Howard, A. W., & Marcy, G. W. 2013,
Proceedings of the National Academy of Science, 110, 19273
Petigura, E. A., Howard, A. W., Lopez, E. D., et al. 2016, ApJ,
Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus,
Rasmussen, C. E., & Williams, C. K. I. 2005, Gaussian Processes
for Machine Learning (Adaptive Computation and Machine
Learning) (The MIT Press)
Rogers, L. A. 2015, ApJ, 801, 41
Schwarz, G. 1978, Annals of Statistics, 6, 461
Sinukoff, E., Howard, A. W., Petigura, E. A., et al. 2016, ApJ,
818, 36
124, 62
827, 78
966
56
Thorngren, D. P., Fortney, J. J., Murray-Clay, R. A., & Lopez,
E. D. 2016, ApJ, 831, 64
Tittemore, W. C., & Wisdom, J. 1989, Icarus, 78, 63
-. 1990, Icarus, 85, 394
Toomre, A. 1964, ApJ, 139, 1217
Valenti, J. A., Butler, R. P., & Marcy, G. W. 1995, PASP, 107,
Van Der Walt, S., Colbert, S. C., & Varoquaux, G. 2011, ArXiv
e-prints, arXiv:1102.1523
Van Eylen, V., Nowak, G., Albrecht, S., et al. 2016a, ApJ, 820,
Van Eylen, V., Albrecht, S., Gandolfi, D., et al. 2016b, ArXiv
e-prints, arXiv:1605.09180
Vanderburg, A., & Johnson, J. A. 2014, PASP, 126, 948
Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, 2198, 362
Weiss, L. M., & Marcy, G. W. 2014, ApJL, 783, L6
−0.2−0.10.00.10.20.30.40.5[Fe/H]31030100MP (Earth-masses)−0.2−0.10.00.10.20.30.40.5[Fe/H]0.00.20.40.60.81.0Menv /MPe > 0.1e < 0.1e uncert.−0.2−0.10.00.10.20.30.40.5[Fe/H]0.3131030100Mcore (Earth-masses)−0.2−0.10.00.10.20.30.40.5[Fe/H]0.3131030100Menv (Earth-masses)24
Figure 14. The bulk heavy element enrichment of the sub-Saturns relative to their parent stars vs. the planet to star mass ratio.
Planets are color-coded by their envelope mass fractions. The dotted line corresponds to the correlation found in Thorngren
et al. (2016) for massive planets MP ≈ 30 − 3000 M⊕.
25
Figure 15. Panels a–d show the planet mass (MP ), core mass fraction (Mcore/MP ), core mass (Mcore), and envelope mass
(Menv) as a function of the number of detected planets in the system. The X-coordinates of each planet has been offset for
legibility. The most massive sub-Saturns tend reside in low-multiplicity systems.
01234567Number of Planets31030100MP (Earth-masses)01234567Number of Planets0.00.20.40.60.81.0Menv /MPe > 0.1e < 0.1e uncert.01234567Number of Planets0.3131030100Mcore (Earth-masses)01234567Number of Planets0.3131030100Menv (Earth-masses) |
0904.4733 | 1 | 0904 | 2009-04-30T02:37:17 | 2006 Fragmentation of Comet 73P/Schwassmann-Wachmann 3B Observed with Subaru/Suprime-Cam | [
"astro-ph.EP"
] | We analyzed the Subaru/Suprime-Cam images of 73P/Schwassmann-Wachmann 3B and detected no fewer than 154 mini-comets. We applied synchrone-syndyne analysis, modified for rocket effect analysis, to the mini-fragment spatial distribution. We found that most of these mini-comets were ejected from fragment B by an outburst occurring around 1 April 2006. The ratio of the rocket force to solar gravity was 7 to 23 times larger than that exerted on fragment B. No significant color variation was found. We examined the surface brightness profiles of all detected fragments and estimated the sizes of 154 fragments. We found that the radius of these mini-fragments was in the 5- to 108-m range (equivalent size of Tunguska impactor). The power-law index of the differential size distribution was q = -3.34 +/- 0.05. Based on this size distribution, we found that about 1-10% of the mass of fragment B was lost in the April 2006 outbursts. Modeling the cometary fragment dynamics revealed that it is likely that mini-fragments smaller than ~10-20 m could be depleted in water ice and become inactive, implying that decameter-sized comet fragments could survive against melting and remain as near-Earth objects. We attempted to detect the dust trail, which was clearly found in infrared wavelengths by Spitzer. No brightness enhancement brighter than 30.0 mag arcsec^-2 (3sigma) was detected in the orbit of fragment B. | astro-ph.EP | astro-ph | 2006 Fragmentation of Comet 73P/Schwassmann-Wachmann
3B Observed with Subaru/Suprime-Cam
Masateru ISHIGURO
National Astronomical Observatory of Japan,
Osawa 2-21-1, Mitaka, Tokyo 181-8588, Japan
[email protected]
Fumihiko USUI
Japan Aerospace Exploration Agency (JAXA), 3-1-1 Yoshinodai, Sagamihara,
Kanagawa 229-8510, JAPAN
Yuki SARUGAKU
Kiso Observatory, Institute of Astronomy, University of Tokyo,
Mitake 10762-30, Kiso, Nagano 397-0101, JAPAN
Munetaka UENO
Japan Aerospace Exploration Agency (JAXA), 3-1-1 Yoshinodai, Sagamihara,
Kanagawa 229-8510, JAPAN
To appears Icarus
Proposed Running head: 2006 Fragmentation of 73P/Schwassmann-Wachmann 3B
Please send communications, proofs and offprint requests to:
Masateru Ishiguro, Dr.
National Astronomical Observatory of Japan,
Osawa 2-21-1, Mitaka, Tokyo 181-8588. Japan
E-mail: [email protected]
2
ABSTRACT
The fragmentation of the split comet 73P/Schwassmann-Wachmann 3 B was observed
with the prime-focus camera Suprime-Cam attached to the Subaru 8.2-m telescope. The
fragmentation revealed dozens of miniature comets (Fuse et al. 2007). We analyzed the
Subaru/Suprime-Cam images, detecting no fewer than 154 mini-comets, mostly
extending to the southwest. Three were close to the projected orbit of fragment B. We
applied synchrone–syndyne analysis, modified for rocket effect analysis, to the
mini-fragment spatial distribution. We found that most of these mini-comets were
ejected from fragment B by an outburst occurring around 1 April 2006, and three
fragments on the leading side of nucleus B could have been released sunward on the
previous return. Several fragments might have been released by successive outbursts
around 24 April and 2 May 2006. The ratio of the rocket force to solar gravity was 7 to
23 times larger than that exerted on fragment B. No significant color variation was
found. The mean color index, V - R = 0.50 ± 0.07, was slightly redder than that of the
Sun and similar to that of the largest fragment, C, which suggests that these
mini-fragments were detected mainly through sunlight reflected by dust particles and
materials on the nuclei. We examined the surface brightness profiles of all detected
fragments and estimated the sizes of 154 fragments. We found that the radius of these
mini-fragments was in the 5- to 108-m range (equivalent size of Tunguska impactor).
The power-law index of the differential size distribution was q = –3.34 ± 0.05. Based on
this size distribution, we found that about 1–10% of the mass of fragment B was lost in
the April 2006 outbursts. Modeling the cometary fragment dynamics (Desvoivres et al.
1999, 2000) revealed that it is likely that mini-fragments smaller than ~10−20 m could
be depleted in water ice and become inactive, implying that decameter-sized comet
fragments could survive against melting and remain as near-Earth objects. We
attempted to detect the dust trail, which was clearly found in infrared wavelengths by
Spitzer. No brightness enhancement brighter than 30.0 mag arcsec–2 (3σ) was detected
in the orbit of fragment B.
Keywords: COMETS, DYNAMICS- COMETS, NUCLEUS- NEAR-EARTH
OBJECTS, INTERPLANETARY MEDIUM
3
1. Introduction
the
is a member of
(hereafter 73P/S-W3)
73P/Schwassmann-Wachmann 3
Jupiter-family comets (JFCs), orbiting the Sun with a 5.4-year period. During the
apparition of 1995, 73P/S-W3 showed a huge outburst in activity. Afterward, four
separate nuclei were confirmed and labeled A, B, C, and D. Of the four, fragment C was
the largest and the presumed principal remnant of the original nucleus. The size of the
nucleus was studied based on the standard assumption for a geometric albedo of 0.04
and a linear phase coefficient of 0.04 mag deg–1; the upper limit of the pre-breakup
radius was 1.1 km (Boehnhardt et al. 1999), and the radius of fragment C was 0.68 ±
0.04 km. Although the radius of fragment B was estimated to be 0.68 km from Hubble
Space Telescope (HST) observations (Toth et al. 2003), Boehnhardt et al. (2002)
established an upper limit of 0.2–0.3 km. Due to poor observing conditions, fragments
A and D were not found in the 2001 apparition.
We had a good opportunity to observe these broken comet fragments during its 2006
return. From near-infrared spectroscopy, no remarkable differences between fragment B
and fragment C were found (Kobayashi et al. 2007; Villanueva et al. 2006). HST
photographed two fragments, B and G, on 18–20 April 2006. These images revealed
several dozen mini-fragments. The Spitzer Space Telescope showed not only many
fragments distributed nearly on orbit but also the debris trail between them. The debris
trail (dust trail) is composed of large dust particles ejected before the last perihelion
passage (Vaubaillon and Reach 2006; Reach et al. 2007). Fuse et al. (2007) made
optical observations of fragment B on 3 May 2006 using the wide-field optical camera
attached to the Subaru 8.2-m telescope. R-band images confirmed 58 mini-comets in the
vicinity of fragment B. No fragments were found along the orbit of fragment B in their
Subaru images (Fuse et al. 2007).
This spectacular Subaru image presents several concerns. We noticed that most of these
fragments were distributed between the anti-solar direction from fragment B and the
negative orbit velocity vector. This positioning was quite interesting because these
mini-comets behaved dynamically like dust particles pressed back by solar radiation
4
pressure against the solar gravity. No obvious dust trail was found in the Subaru optical
image, even though it was clear in the Spitzer infrared image. In this study, we
re-analyzed the Subaru/Suprime-Cam images using the masking method developed for
the detection of faint cometary dust clouds (Ishiguro et al. 2007; Sarugaku et al. 2007;
Ishiguro 2008) and constructed a comet image without contaminants (e.g., stars and
galaxies). This technique enabled us to detect mini-comets brighter that 26.5 mag and
diffuse light sources associated with the comet brighter than 30.0 mag arcsec–2.
Applying a unique method of examining fragment size and onset time (modified
synchrones and syndynes), we examined the dynamical properties of the mini-fragments.
We also studied the brightness profile of these mini-comets and deduced their sizes.
Given the dynamical properties and sizes, we considered the activity of the
mini-comets.
5
2. Data and Observations
2.1. Data
We re-analyzed the Subaru data provided by the SMOKA data server, which is operated
by the Astronomy Data Center, National Astronomical Observatory of Japan (Baba et al.
2002). Observations of 73P/S-W3 were carried out by Fuse et al. (2007) using the
Subaru 8.2-m telescope on Mauna Kea, Hawaii, on a single day, 3 May 2006, when
fragment B was at a heliocentric distance rh = 1.070 AU, a geocentric distance Δ =
0.112 AU, and a solar phase angle α = 54°. Fuse et al. (2007) used an optical CCD
camera, Suprime-Cam, attached to the prime focus of Subaru. This combination
provided wide-field imaging capability, 34’ × 27’, with a pixel resolution of 0.20”
pixel–1. The seeing was about 0.7” (FWHM), which projects to 57 km at the position of
the comet. The exposure time and filters are summarized in Table 1. All comet images
were taken in comet-tracking mode. Although Fuse et al. (2007) did not use the short
exposure-time R-band images (10−30 s) and V-band images, we found that these were
essential to (i) determining the brightness of mini-comets near fragment B, (ii)
improving the signal-to-noise ratio, and (iii) identifying detected sources as having a
cometary origin. Fragment B was so bright that a large area of sky near B was saturated
in 120-s exposures. The signal-to-noise ratio was improved by combining all images.
We used the V and R composite image to confirm the mini-fragments because the color
index V-R avoids false detections (see Section 2.2). Further explanations of the
Suprime-Cam and 73P/S-W3B observations appear in Miyazaki et al. (2002) and Fuse
et al. (2007), respectively.
[Table 1]
2.2. Data Reduction
As a first step, the obtained data were reduced in the standard way with bias and
flat-field corrections. These ancillary data were also provided through the SMOKA
6
system. Because useful bias data were not obtained during the night of 3 May, we used
bias frames taken on 1 May. Flux calibration was done using Landolt standard stars in
the SA113 region (Landolt 1992).
The sky background was contaminated by elongated stars and galaxies because the
observations were carried out in comet-tracking mode. We removed these stellar objects
using the masking algorism developed for the data reduction of cometary dust trails
(Sarugaku et al. 2007; Ishiguro et al. 2007; Ishiguro 2008), outlined as follows. We first
made images to align the stars, because this is an effective way of detecting faint stars
and galaxies. Using these images, stars were automatically detected by a source
extractor program, SExtractor (Bertin and Arnouts 1996). We masked the identified
objects using 18” × 6” rectangular masks. We also masked pixels identified as bad in
the bias (hot pixels and lines) and flat-fielding images (pixels with sensitivity 10%
higher or lower than the average). We combined the masked images with offsets to
align the comet, excluding the masked pixels and shifting the background intensity to
zero. Because the comet moved relative to the stars, it was possible to exclude nearly all
masked pixels in the resultant composite image. Finally, we obtained V- and R-band
composite images without stars. For the composite images, we used images with 120-s
exposure times. Approximately 17% of the pixel values in the images were masked by
this method. Therefore, the effective total exposure times were 400 s in each
wavelength.
To extract the mini-comets in the composite images, we first flattened the sky
background by subtracting the 23-pixel × 23-pixel (4.6” × 4.6”) running median images.
This is a standard image-processing technique known as “unsharp masking.” The
large-scale components associated with the mini-comets could be subtracted out by this
method. The R-band image is shown in Fig. 1. Because we detected no significant
difference in appearance between the V- and R-band subtracted images, we combined
these two into single image. This VR composite image was used for the detection of
faint mini-comets. We used the SExtractor again to detect the mini-comets. We found
211 mini-comet candidates in this VR composite image. The positions and magnitudes
of these mini-comets were examined using the “phot” command in the IRAF/APPHOT
package. We set a fixed aperture size of 2.0”. This aperture gathered the light from
7
nuclei and a portion of the light from the coma components. Assuming that fragment B
was at the brightest point in the 10 s exposure image, we determined the relative
positions of the mini-comets.
[Figure 1]
Of the 211 mini-comet candidates, we determined R-band magnitudes for 176 objects,
V-band magnitudes for 161 objects, and both R- and V-band magnitudes for 154 objects.
In Fig. 2, we compare the V- and R-band magnitudes of 154 comet candidates. A glance
at Fig. 2 reveals that the V-R color indices of these 154 objects were slightly redder that
of the Sun (V-RSun = 0.367; Rabinowitz 1998). The mean color of the mini-comets, V-R
= 0.50 ± 0.07, was similar to that of main nucleus C, V-R = 0.48 ± 0.17 (Boehnhardt et
al. 1999; Lamy et al. 2004). Accordingly, we can state that at least 154 mini-comets
were detected by our data analysis methods.
[Figure 2]
In Fig. 1(b), we find that some mini-fragments were elongated in the anti-solar direction.
Fragment B was also elongated anti-sunward. Because these dust particles were
strongly coupled to solar radiation, they could have been small particles or highly
porous dust aggregates (Mukai et al. 1992; Kimura et al. 2002, 2003). We also found a
disconnection near fragment B (see Fig. 1(c)). In general, this disconnection could have
resulted from recent fragmentation (discussed below), accidental eruption of dust
particles, or solar magnetic field reversal.
3. Results and Discussion
3.1. Observed Mini-Fragment Spatial Distribution
Figure 3 shows the position of 154 mini-comets relative to that of fragment B. As we
8
described in Section 1, most of fragments were distributed toward the southwest (the
lower right of Fig. 1). Two or three objects appeared on the trailing side of B and close
to its projected orbit. Three objects appeared on the leading side of B (Fig. 1(d)). It
appears that the three objects on the leading side were released with a sunward velocity
component on the previous return, giving these three fragments a smaller semi-major
axis than that of B, whereas most of the objects in the southwest were ejected at the
current apparition and expanded by the rocket effect. Figure 4 is a histogram of the
position angle of the mini-fragments, which is defined as an angular offset of the
mini-comet to fragment B relative to the north celestial pole. East, south, and west
correspond to position angles of 90°, 180°, and 270°, respectively. The southwest
population is distributed nearly symmetrically with an average of 237.7°. A small
peak appears around 217.5°, which we discuss in Section 3.3. In Fig. 5, we show the
R-band magnitudes of these southwest comets with respect to the distance between B
and each mini-comet. As a general trend, the bright mini-comets were located near B,
whereas faint objects were distributed far from fragment B.
[Figure 3]
[Figure 4]
[Figure 5]
3.2. Interpretation: Basic Dynamics Equations
Let us consider the observed distribution of the mini-fragments from the standpoint of
dynamics. The motions of “dry dust particles” and “icy comets” should differ from one
another. The motion of icy comets, composed of refractory (silicates and CHON
particles) and volatile (mainly H2O) particles, is governed primarily by solar gravity and
perturbed by the planets’ gravities. When these mini-comets were in the inner solar
system, the rocket effect, which induces a recoil force from the gas outflow momentum,
could continuously perturb their orbits. Although the recoil force is generally referred to
as the “non-gravitational force,” we use the term “rocket force” in this paper to
distinguish it from other non-gravitational forces, such as radiation pressure.
9
The equation of motion of the mini-fragments in the rectangular coordinate system can
be written as
GM!m c
2
rh
e r +
N p
#
i=1
GM im c
2
ri
. (1)
(
e Pi + m c F1e r + F2e T + F3e N
)
Fc = "
Here G and M are the gravitational constant and the mass of the Sun, respectively; rh is
the distance between the Sun and comet; and mc is the mass of the mini-fragment. The
second term on the right side of Eq. 1 denotes the planetary perturbations; Mi is the
mass of the i-th planet, and ri is the distance between the i-th planet and the comet. We
considered 10 objects (Np = 10: eight planets, Pluto, and the Moon). We used DE406
provided by NASA/JPL for the ephemerides of these 10 objects. ePi is the comet–planet
unit vector. F1, F2, and F3 represent the acceleration by the rocket effect: F1 is the
acceleration along the radial vector defined outward along the Sun–comet line; F2,
perpendicular to the radial vector in the orbit plane and toward the comet’s direction of
motion; and F3, perpendicular to the orbit plane. er, eT, and eN are the three unit vectors
along the directions of the three rocket forces that satisfy the condition eN = er × eT. The
acceleration components from the rocket effect can be considered a function of
heliocentric distance, conventionally written
F j rh(
where g(rh) expresses the water ice vaporization rate as a function of heliocentric
distance rh
) , (2)
) = A j g rh(
g rh(
) = "TII
#
rh
%
rTII
$
) m TII
&
(
’
*
,
1 +
,
+
#
rh
%
rTII
$
n TII
&
(
’
) kTII
-
/
/
.
, (3)
where rTII = 2.808 AU, mTII = 2.15, nTII = 5.093, and kTII = 4.6142. The value of αΤΙΙ is
chosen such that g(rh = 1) = 1, which gives αΤΙΙ = 0.111262. Aj in Eq. 2 is referred to
as the “Type-II non-gravitational parameter” and widely applied to describe the
non-gravitational motion of comets due to rocket effect (Marsden et al. 1973). The
10
!
!
!
transverse component parameter A2 is almost always well determined for short periodic
comets because it is sensitive to the secular change in the semi-major axis, which is
established using long observation intervals. A1 is sensitive to perturbations of the
longitude of perihelion and is often not as well determined. A3 is usually the least
well-determined of the three because it is sensitive to perturbations in the orbital
inclination and longitude of the ascending node and neither of these perturbations is
secular (Yeomans et al. 2005).
In contrast, the orbit of dust particles is determined by the solar radiation pressure,
Poynting-Robertson drag, and so forth, as well as gravitational forces. The dust particles
may not include ice components because of their short lifetime. Mukai (1986) studied
the lifetime of water ice and found the lifetime for 1-mm particles to be less than a day
at 1 AU. Ignoring the rocket force, we can express dust particle motion as
-
’
)
/ +
(
.
v
c
N p
0
i=1
GM imd
2
ri
GM!md
2
rh
*
$
r
1 " #(
)e r " #
&
e r +
,
%
+
c
e Pi , (4)
Fd = "
where β is the ratio of the solar radiation pressure with respect to the solar gravity and v
is the orbital velocity of the dust particle. The first term in the bracket comes from the
solar gravity reduced by the radiation pressure, whereas the second term is derived from
the Poynting-Robertson drag (Burns et al. 1979). In addition to these forces from Eqs. 1
and Eq. 4, the solar wind drag and the Yarkovsky effect may perturb the orbits of dust
particles and cometary fragments. As studied by Mukai and Yamamoto (1982), solar
wind drag is not efficient for particles larger than 1 µm. The Yarkovsky effect is also
ineffective for the short-term evolution under discussion here (less than several years).
Thus, we ignored the solar wind drag and the Yarkovsky effect.
For a spherical particle of radius a (cm) and mass density ρ (kg m–3), β is defined as
!
" =
KQpr
#a
, (5)
where K is the constant
!
11
, (6)
3L!
K =
16"GM!
L and c are the solar luminosity and speed of light, respectively. Qpr, which is
defined as Qpr = Qext—<cosθ>Qsca, is the radiation pressure coefficient averaged over
the solar spectrum. Here, Qext and Qsca are the efficiency factors for extinction and
scattering, respectively, and <cosθ> is the asymmetry parameter for light scattering. Qpr
is the radiation pressure coefficient averaged over the solar spectrum (Burns et al. 1979).
Assuming that the particles are compact in shape and large compared to the optical
wavelength (>>0.5µm), we can fix Qpr = 1. From observations, it is known that
Jupiter-family comets eject dust particles with β = 6 × 10–6 – 0.2 (Fulle 2004). It is
inferred that comet brightness may be dominated by light scattered by the largest
particles (i.e., smallest β; Fulle 2004, Ishiguro et al. 2007).
3.3. Interpreting Spatial Distribution using Modified Synchrones and Syndynes
We assumed that dust and gas emission occurred on the sunlit hemisphere of each
fragment, symmetric with respect to the comet–Sun axis. This assumption is reasonable
because large portions of these mini-fragments were covered with fresh icy surface
when they were produced. We then expected that the rocket effect was exerted in the
anti-solar direction alone. In fact, A1 of fragment B is one order of magnitude larger
than A2 and A3 (Sekanina 2005), which supports this assumption. Here, we define the
ratio of rocket force acceleration with respect to the solar gravitational acceleration as
+
-
"rkt rh(
,
-
F2 = F3 = 0
.
where βrkt,0 is the ratio at rh = 1 AU. The position of the mini-comet parameterized by
βrkt,0 can be obtained by solving the following equation:
, (7)
# "rkt ,0 g rh(
2
)rh
) # F1 rh(
$
GM!
)
&
2
rh
%
*1
’
)
(
12
!
!
!
N p
$
i=1
GM im c
2
ri
e Pi
GM!m c
2
rh
)
]e r +
[
1 " #rkt rh(
. (8)
Fc = "
At small heliocentric distances (rh < 1.5 AU), we can assume that βrkt is independent of
the heliocentric distance because F1 (which is proportional to the water sublimation
2)
rate; see Eqs. 2 and 3) is approximately proportional to the solar irradiation (i.e., ∝rh
at rh < 1.5 AU. The mathematical form of the rocket force F1 at rh < 1.5 AU is similar to
that of the solar radiation pressure term of the dust particle in Eq. 4. Accordingly, we
can apply the synchrones and syndynes analysis for a small heliocentric distance. Figure
6 compares the synchrones and syndynes of dust particles (obtained by Eq. 4) with
those of mini-comets (obtained by Eq. 5). A significant difference in the synchrones
appeared when the fragment was ejected at rh > 2 AU for the above reason.
[Figure 6]
The notion of synchrones and syndynes was originally introduced to fit the observed
dust tail (see e.g., Finson and Probstein 1968). In this paper we use the term “modified
synchrones and syndynes” to refer to the locus of mini-fragments. In Fig. 7, we
compare the positions of mini-comets and the modified synchrones and syndynes. The
positions of the mini-fragments are illustrated by cross signs, and the modified
synchrones and syndynes by dashed and solid lines, respectively. Most of fragments
were concentrated in the range of βrkt = 3 × 10–4 and βrkt = 1 × 10–3. This value is about
7–23 times larger than that exerted on fragment B.
[Figure 7]
An advantage of using the modified synchrones and syndynes is that we are able to
determine the onset time of fragmentation using a “snapshot.” In Fig. 4, we find a
prominent concentration at the position angle 237.5°. This position angle coincides with
the synchrone of 26 March 2006. What happened on that date? Figure 8 shows the light
curve of fragment B (obtained from Seiichi Yoshida’s web site, http://www.aerith.net/).
The plotted magnitude HΔ was normalized to the geocentric distance Δ = 1 AU. Two
arrows, labeled O/B and SU, denote the time of expected onset on 26 March and the
13
time of the Subaru/Suprime-Cam observation, respectively. In Fig. 8, a brightness
enhancement appeared a few days after 26 March. Therefore, we can argue that most of
mini-comets in the southwest were released by an outburst occurring in late March or
early April. The onset time of 26 March is earlier than the generally described outburst
time of 1 April (Sekanina 2007). This small discrepancy may imply that these
mini-fragments were active and progressive during the early stage of the ejection and
became inactive over a one-month period. Similar evidence was found for the broken
comet C/1999 S4 (LINEAR): Weaver et al. (2001) estimated the separation time from
the dynamical properties of 100-m mini-fragments and found that their results indicated
an earlier time than the commonly accepted disruption time of C/LINEAR. In addition
to the outburst on 1 April, Sekanina (2007) argued that successive outbursts occurred on
24 April and 2 May, although, in Fig. 8, we cannot find evidence for an outburst on 24
April. It is likely that the detached feature in Fig. 1(c) was the fragment produced
around 2 May, and five fragments at a position angle of around 217.5° (see also the 24
April synchrone in Fig. 7(a)) might have been ejected by the outburst on 24 April (the
synchrone of 24 April coincides with a position angle of 219°). It should be emphasized,
however, that the outstanding single peak at a position angle of 235.5° in Fig. 4
indicates that most of the fragments in the southwest were released by the 1 April
outburst.
[Figure 8]
We applied a model for the dynamics of cometary fragments to our data, following
Desvoivres et al. (1999, 2000). They considered the energy balance on the surface of the
icy body, given by
S0
2
rh
(
1 " pv
) cos z = #$T 4 + Lw T( )
dZ
dt
, (9)
where S0 is the solar flux at 1 AU, pv is the geometric albedo in V-band, z is the zenith
distance of the Sun, ε is the emissivity, and σ is the Stefan–Boltzmann constant. T
denotes the equilibrium temperature. Note that heat transferred to the deeper layers is
neglected in this model (Desvoivres et al. 1999, 2000). The latent heat of sublimation of
14
!
the water, Lw, is given by (Delsemme & Miller 1971)
Lw T( ) = 2.886 " 10 6 #1116
T
$
&
%
1K
’
) J kg#1 . (10)
(
The sublimation rate of the water ice is given by
dZ
dt
=
1
1 + 1 "
# T( )Pw T( )
mw
2$kT
kg s-1 , (11)
–1 (ρw and ρd are the
where κ is the water ice-to-dust mass ratio, defined as κ = ρw ρd
masses of water ice and dust particles per unit volume, respectively). mw is the
molecular mass of the water. k is the Boltzmann constant. γ denotes the dimensionless
sticking coefficient (Haynes et al. 1992; Enzian et al. 1997) given by
" T( ) = #2.1 $ 10
%
#3 T
’
&
1K
(
* +1.042
)
(
T > 20K
) . (12)
the saturated vapor pressure of water Pw(T)
In Eq. 11,
Clausius-Clapeyron equation:
is given by
the
*
$
Pw T( ) = 3.56 " 1012 exp #6141.667
,
&
%
1K
+
T
#1
’
)
(
-
/
.
Pa . (13)
Assuming a spherical body, the anti-solar acceleration due to sublimation of ice is given
by an integral over the sunlit hemisphere:
F1 =
2
"fRc
m c
"
2$
o
dZ
dt
dz , (14)
v th#sin 2 z
where Rc is the radius of mini-comet, φ is the geometric correction factor, and vth is the
mean velocity of sublimating water given by
15
!
!
!
!
!
v th =
8kT
"mw
. (15)
We introduced the fractional active area f into Desvoivres’s original model. f = 1
indicates no dust accumulation on the entire surface area. We rewrote the first equation
of Eq. 7 as follows:
2 f
3rh
v th sin 2 z
dz
dZ
#
2$
o
dt
4GM!Rc%c
, (16)
"rkt =
where ρc is the mass density of the mini-comet, satisfying the condition ρc = ρw + ρd.
From Eq. 16, it is clear that βrkt is inversely proportional to Rc and ρc, and proportional
to f. Equation 16 suggests that smaller fragments were accelerated to higher velocities
relative to fragment B, whereas larger fragments were accelerated to lower velocities,
qualitatively supporting the result that the brightness of the mini-fragments decreased as
the distance from fragment B increased (Fig. 5).
As discussed in Desvoivres et al. (1999), the non-gravitational acceleration by the
rocket effect is not sensitive to κ and pv. We thus used κ = 1 and pv = 0.04. We assumed
ε = 0.9 and φ = 2/3, following Desvoivres et al. (1999). βrkt of the fragment B around
1995 (when the initial outburst occurred) was estimated to be 4.4 × 10–5 (Sekanina
1996), and the radius of fragment B is 680 m (Toth et al. 2003) or <~300 m
(Boehnhardt et al. 2002). By substituting Rc = 680 m or ~300 m and βrkt = 4.4 × 10–5
into Eq. 16 we obtain
*
$
"c
,
&
1000 kg m#3
,
%
+
$
"c
,
&
,
1000 kg m#3
%
-
for fragment B during the 1995 apparition. Given a mass density ρc = 200–800 kg m–3
(Sosa and Fernandez 2008), we found f > 0.5 (when Rc = 680 m) or 0.2 < f < 0.9 (when
(17)
#1
’
)
(
#1
’
)
(
f = 1.1
Rc = 300m
f = 2.5
Rc = 680m
16
!
!
!
Rc = 300 m), indicating that a large fraction of fragment B’s surface was very active
shortly after birth. It is reasonable to think that a large portion of the icy surface on
fragment B was exposed in the 1995 breakup. Applying Eq. (16), we can calculate a
typical mini-comet radius (βrkt = 3 × 10–4 to 1 × 10–3) of about 10–100 m, assuming that
f and ρc of the mini-comets were the same as those of fragment B. This size estimate is
roughly consistent with the results obtained by the photometry in Section 3.4. Further
discussion on the size and the activity will occur in the next section.
3.4. Photometric Results
It is possible to derive the size of mini-comets from photometry. The magnitudes we
obtained from aperture photometry in Section 2.2 must overestimate the brightness of
the nuclei because of the effects of the near-nucleus coma. The rocket effect suggests
the existence of sublimating ices. To examine the non-stellar nature of mini-comets, we
constructed normalized radial surface brightness profiles for the mini-comets and
compared them to stellar profiles obtained with sidereal tracking. The stellar profile was
constructed using the exposure that was taken to bring the camera into focus. We found
that the stellar profile was similar to the one-dimensional surface brightness profile of
the field stars in the comet exposures, suggesting little variation in seeing over the
observation period. Figure 9 shows example brightness profiles for the mini-comets and
the reference star. In the graph, we show two extreme cases: “mini-comet 1,” apparently
the most active; and “mini-comet 3,” inactive. The brightness profiles of the other
mini-comets are intermediate between these two (“mini-comet 2” is the example). To
deduce the size of the mini-comets, we adapted the method of Lamy et al. (1998).
Assuming that the coma surface brightness decreases as ρ-η, where ρ is the projected
distance between the line of sight and the nucleus, we estimated the flux from the
nucleus. The corresponding total surface brightness distribution, B(ρ), is thus given by
[
B "( ) = kcoma"#$ + k nucleus% "( )
where δ(ρ) is the Dirac δ function and ⊗ is the convolution operator. PSF is the point
spread function obtained by the reference stars as mentioned above. kcoma and knucleus are
the brightness scaling factors of the coma and the nucleus, respectively. Because the
] & PSF , (18)
17
!
!
coma brightness of the small fragment is too faint to determine from its surface
brightness, we used the typical value of the large fragment, i.e., η = –2. Using this
procedure, we found that the comet’s nuclei contribute >25% of the total scattering
cross-section measured with a 2” aperture.
[Figure 9]
The apparent R-band magnitude of the airless body in the solar system is written
following Fernandez et al. (2000) and Jewitt (2006) as
{
2
5
#
m R # b$# m!
} , (19)
2"2
2 = Crh
pR RC
10
where C is the constant value 2.25 × 1022 m2, mR is the R-band magnitude of the
fragment’s nucleus at the heliocentric distance rh and the geocentric distance Δ, and m
is the solar R-magnitude (–27.1). The R-band albedo pR is in the range pR ~ pV =
0.02–0.08 (Lamy et al. 2004; Tancredi 2006). Because the data were taken at the phase
angle α = 54°, the observed magnitude was highly influenced by the scattering phase
function. We correct the phase effect by the linear law. The slope b of the phase
function is in the range 0.02–0.06 mag deg–1 (Lamy et al. 2004; Fernandez et al. 2000).
Given that pR = 0.04 and b = 0.04 mag deg–1, as is typical of the nuclei of short-period
comets (Fernandez et al. 2005), we obtained the size of the fragment in the range of Rc
= 5–108 m. This is consistent with the size estimate by βrkt in Section 3.2. Uncertainty
in the phase function caused a 60% error in Rc, whereas a 50% error in the geometric
albedo translated into a 25% error in Rc. The uncertainty in the correction for coma
contamination is at most 20%. From these uncertainties, we consider that the radius of
the bare nucleus is uncertain by a factor of ~2 or less.
3.5. Total Mass Loss and Brightness Increase
Given that all fragments have the same albedo value and phase function slope, we are
18
able to deduce the size distribution. The cumulative size distribution is shown in Fig. 10.
Filled circles denote the size determined by standard assumptions for the geometric
albedo and the linear phase coefficient, i.e., pR = 0.04 and b = 0.04 mag deg–1.
[Figure 10]
Because large fragments have a stochastic problem and small fragments have large
measurement uncertainties, we fit the slope to the cumulative fragment size distribution
between 12 m and 37 m. We found that the differential size distribution, defined as
N Rc(
)dRc = N ref
"
Rc
$ $
Rref
#
q
%
’ ’
&
dRc , (20)
was q = –3.34 ± 0.05. The number of fragments with a reference size Rref = 1 m is also
obtained from the above fit, and found to be Nref = (3.40 ± 0.20) × 104. The power-law
index of size distribution turns out to be similar to the index of the theoretical
distribution, q = –3.5, expect for self-similar collision cascades as predicted by
Dohnanyi (1969), but steeper than the slope of short-period comets (q = 2.4–2.6; Meech
et al. 2004, Weissman and Lowry 2003). It is interesting that the slope is quite similar to
that of the dust particles in the dust trails of 2P/Encke, 4P/Faye, 22P/Kopff,
67P/Churyumov-Gerasimenko (Ishiguro et al. 2007; Sarugaku et al. 2007; Ishiguro
2008), and dust tails (Fulle 2004; Fulle 1992; Fulle 1990). This similarity suggests that
the cometary debris size distribution might be a simple power law distribution over a
wide size range. Because q > –4, the total mass of the fragments Mtotal strongly depends
on the largest fragment. This enables us to deduce the total mass accurately. The total
mass Mtotal is given by
M total =
)
,
(
N > R ref
.
%
"#
.
-
i=1
4
3
3 +
Ri
$
R ref
R min
3
N ref Rref
&
R
( (
Rref
’
3 + q
)
+ +
*
/
. (21)
1
dR
1
0
!
!
19
Using the parameter above, the total mass of the fragments was 1.62 × 1010 (ρc/103 kg
m–3) kg for a standard geometric albedo and linear phase coefficient. Note that there
should be a factor of 8 or less uncertainty due to the unknown albedo and the linear
phase coefficient. Applying the Rc = 300−680 m of fragment B [equivalent to a mass of
0.11−1.32 × 1012 (ρc (103 kg m–3)–1 ) kg], we found that 1−10% of the mass was lost in
the outbursts of April 2006.
We next considered the brightness change on 1 April. The magnitude of fragment B
increased by ΔHΔ ~ 3 from HΔ ~ 14.5 to HΔ ~ 11.5 around 1 April (see Fig. 8). The total
cross-section of the material ejected by the outburst depends on the smallest particles
because q < –3. The cross section is given by
)
+
(
N > R ref
-
$
C total = "
-
,
i=1
2 +
Ri
#
R ref
R min
N ref
%
R
’ ’
Rref
&
2 + q
(
* *
)
.
0
dR
0
/
. (22)
1 2
N ref
3 + q
%
Rmin
’ ’
Rref
&
3 + q
(
* *
)
(
q < 23
)
Particles with Rmin ~< λ * 2π−1 will scatter optical light inefficiently because of
Rayleigh scattering. Therefore, we set the minimum size Rmin = 1.0 × 10–7 m (0.1 µm).
If we assume that the scattering properties (i.e., albedo and phase function) of a
sub-micron particle are the same as those of the decameter-sized fragments, we can
deduce the brightness increase by fragmentation using Eqs. 19 and 22. We found an
estimated magnitude increase of 1.3 ± 0.6 mag. Admittedly, the calculated flux increase
is less than half the observed flux increase; nonetheless, we argue that our estimate for
the power-law index is appropriate for explaining the observations because (i) the other
light sources we considered in Eq. (21), such as gas emission from volatiles and
scattered sunlight by secondary dust particles released from the fragments, must have
been present; and (ii) the scattering properties of sub-micron dust particles contain
uncertainties. It might be better to state that Eq. (21) gives a lower limit of q for
submicron particles; that is, if q < –3.7, the magnitude increase exceeds the observed
increase of ΔHΔ ~ 3 mag. Therefore, it seems reasonable to think that the power-law
20
!
index of size distribution of the comet debris is about 3.3–3.4 over a wide size range.
3.6. Active Area Fraction
As described above, the positions of the mini-comets relative to fragment B result from
acceleration by rocket force. Because the observed separation between B and the
mini-comets is the distance projected on the celestial plane, it is not easy to determine
the real distance. We drew lines perpendicular to the synchrone of 1 April and derived
βrkt of the fragment from the distance between B and the foot of a perpendicular. By
–1. The result is shown in Fig. 11.
applying Eq. 16, we obtained f ρc
[Figure 11]
–1, and the mass density is unknown. Here, assuming that the
The left axis denotes f ρc
mass density of the mini-comets is the same as 9P/Tempel 1 and 81P/Wild 2 (i.e., ρc ~
450 kg m–3), we found that all fragments except one (Rc = 100 m) have a fractional
active area f < 1. Perhaps fragments with f = 1.4 and Rc = 100 m have low densities (ρc
< 320 kg m–3) to fit the f < 1 condition. It is likely that f ρc
–1 decreases with decreasing
radius, becoming almost constant below ≈10–20 m. This suggests that water ice in the
surface layer within <10 m might be exhausted by sublimation. In fact, the
–1, suggesting that this
point-source-like comet in Fig. 9 had the smallest value of f ρc
comet was depleted in water ice. Because small fragments can become inactive even
around 1 AU, we would suggest that such decameter-sized comet fragments could
remain as near-Earth objects. Figure 11 illustrates our presumption that a fraction of
near-Earth asteroids could be produced by fragmentations of comets as seen in
73P/S-W.
3.7. Dust Trail
The appearance of our optical
the
that of
is quite different from
image
infrared-wavelength image taken by Spitzer. The infrared images showed not only the
mini-comets but also the dust trails connecting each fragment (Vaubaillon and Reach
2006; Reach et al. 2007). Figure 12 shows a cut profile of the sky background
21
perpendicular to the comet orbit. The cut profile is the result averaged more than ~20’
along the projected orbit. No signal from the dust trail could be found. We put an upper
limit on the surface brightness of 30.0 mag arcsec–2 (3σ). The upper limit of the surface
brightness is converted into the optical depth multiplied by albedo A(θ) × τ, and we
found that A(θ) × τ = 2.2 × 10–12. A comparative study between optical and infrared
observations of the dust trail would provide information about the optical properties of
the dust trail.
[Figure 12]
4. Summary and Remarks
In this paper, we have described the data analysis and interpretation of the optical image
of 73P/Schwassmann-Wachmann 3B taken by the Subaru/Suprime-Cam. We detected
at least 154 mini-fragments whose colors were similar to those of the Sun. Except for a
few fragments, they were systematically distributed toward the southwest. We applied a
modified synchrone-syndyne analysis to the spatial distribution of a number of
mini-fragments from the standpoint of dynamical evolution by rocket force. We found
that most of these mini-comets were ejected from fragment B on 1 April 2006. This
result is consistent with the evidence that the magnitude of B surged around 1 April.
Three fragments on the leading side of B could have been released with a sunward
velocity component on the previous return. Several fragments at a position angle of
around 217.5° might have been released by an outburst around 24 April, and a detached
feature near B could be related to an outburst around 2 May 2006. The ratio of rocket
force to solar gravity is approximately 7−23 times larger than that exerted on B, which
implies that the radii of the mini-comets are about 10–100 m. No significant color
variation was found; the mean color index V-R = 0.50 ± 0.07 was slightly redder than
the solar color, suggesting that these mini-comets were observable mainly through
sunlight scattered by red dust particles and nuclei surfaces. We examined the brightness
profiles of all detected mini-comets and deduced the sizes of 154 fragments. The
power-law index of the differential size distribution was q = –3.34 ± 0.05. Based on the
size distribution, we found that about 1–10% of fragment B’s mass was lost in outbursts
occurring in 2006. No dust trail was detected along fragment B’s orbit. We placed an
22
upper limit on the surface brightness of 30.0 mag arcsec–2 (3σ) and on the optical depth
multiplied by albedo, A(θ) × τ = 2.2 × 10–12.
–1, including the energy balance on the icy
We applied Desvoivres’s model to derive f ρc
surface and ignoring dust mantle accumulation. Although small dust particles can be
lifted by escaping gas pressure, large dust particles cannot escape and gradually
accumulate on the surface. This thin dust layer may significantly diminish the
vaporization of water (Prialnik and Bar-Nun 1988; Rosenberg and Prialnik 2007;
Prialnik et al. 2008). In fact, the surface temperature on the large portion of 9P/Tempel
1 is in good agreement with the temperature for the standard thermal model, which
suggests that cometary surfaces could be covered with a dust mantle (A’Hearn et al.
2005). It should be noted that, for the model, we included an unrealistic assumption in
Eqs. 9–16 that the mass density (ρc, ρw, and ρd) and the water ice-to-dust mass ratio (κ)
were constant. It is natural to think that ρw and κ near the surface decrease with water
ice sublimation. The dust density near the surface could also decrease because the
sublimating water vapor pushes the dust particles into interplanetary space. The fraction
of active surface area f could decrease due to ice depletion. However, we could not
incorporate the depletion of ice near the surface in the model (Eqs. 9–16). Studying the
vaporization of water from the subsurface layer is part of our future work.
23
Acknowledgement
This Suprime-Cam image was taken using Subaru Telescope, and obtained from the
SMOKA, which is operated by the Astronomy Data Center, National Astronomical
Observatory of Japan. The visible lightcurve data were obtained from Seiichi Yoshida’s
web site. We sincerely thank people who are involved in these projects. We also thank
Dr. T. Fuse (Subaru observatory) and his colleagues, for their effort in the data
acquisition by Subaru. Also we acknowledge two anonymous referees, Prof. J.
Watanabe, and Dr. T. Kasuga for their valuable comments. The ephemeris data, DE406,
was provided by NASA/JPL.
24
A’Hearn M. S., et al. 2005. Deep Impact: Excavating Comet Tempel 1. Science 310,
5746, 258-264.
Baba, H.,asuda, N. Ichikawa, S. Yagi, M., Iwamuro, N., Takata, T. Horaguchi, T., Toga,
M., Watanabe, M., Ozawa, T., Hamabe, M. 2002, Development of
the
Subaru-Mitaka-Okayama-Kiso Archive System. Astronomical Data Analysis Software
and Systems XI, eds. D. A. Bohlender, D. Durand, & T. H. Handley, ASP Conference
Series, 281, 298.
Bertin, E., Arnouts, S. 1996. SExtractor: Software for source extraction. Astron.
Astrophys. Suppl., 117, 393-404.
Boehnhardt, H., Rainer, N., Birkle, K., Schwehm, G. 1999. The nuclei of comets
26P/Grigg-Skjellerup and 73P/Schwassmann-Wachmann 3. Astron. Astrophys. 341,
912-917.
Boehnhardt, H., Holdstock, S., Hainaut, O., Tozzi,G. P., Benetti, S., Licandro, J. 2002.
73p/Schwassmann-Wachmann 3 - One Orbit after Break-Up: Search for Fragments.
Earth, Moon, and Planets 90, 1, 131-139.
Burns, J. A., Lamy, P. L., Soter, S. 1979. Radiation forces on small particles in the solar
system. Icarus 40, 1-48.
Delsemme, A. H. & Miller, D. C. 1971. The continuum of Comet Burnham (1960 II):
The differentiation of a short period comet. Planet. Space Sci. 19, 1229-1257.
Desvoivres, E., Klinger, J., Levasseur-Regourd, A. C., Lecacheux, J., Jorda, L., Enzian,
A., Colas, F., Frappa, E., Laques, P. 1999. Comet C/1996 B2 Hyakutake: observations,
interpretation and modelling of the dynamics of fragments of cometary nuclei. Monthly
Notices of the Royal Astronomical Society 303, 4, 826-834.
Desvoivres, E., Klinger, J., Levasseur-Regourd, A. C., Jones, G. H. 2000. Modeling the
25
Dynamics of Cometary Fragments: Application to Comet C/1996 B2 Hyakutake. Icarus
144, 172-181.
Dohnanyi, J. W. 1969. Collisional models of asteroids and their debris. J. Geophys. Res.
74, 2531-2554.
Enzian, A., Cabot, H., Klinger, J. 1997. A 2 1/2 D thermodynamic model of cometary
nuclei. I. Application to the activity of comet 29P/Schwassmann-Wachmann 1. Astron.
and Astrophys. 319, 995-1006.
Fernández, Y. R., Lisse, C. M., Ulrich K. H., Peschke, S. B., Weaver, H. A., A'Hearn,
M. F., Lamy, P. P. , Livengood, T. A., Kostiuk, T. 2000. Physical Properties of the
Nucleus of Comet 2P/Encke. Icarus 147, 145-160.
Fernández, Y. R., Jewitt, D. C., Sheppard, S. S. 2005. Albedos of Asteroids in
Comet-Like Orbits. Astron. J.130, 1, 308-318.
Finson, M. L., Probstein, R. F. 1968. A theory of dust comets. I. Model and equations
Astrophys. J. 154, 353-380.
Fulle, M. 1990. Meteoroids from short period comets. Astron. Astrophys. 230, 1,
220-226.
Fulle, M. 1992. Dust from short-period Comet P/Schwassmann-Wachmann 1 and
replenishment of the interplanetary dust cloud. Nature 359, 6390, 42-44.
Fulle, M. 2004. Motion of Cometary Dust. In Comets II (M.C. Festou et al., eds),
University of Arizona, Tucson, 565-575.
Fuse T., Yamamoto N., Kinoshita D., Furusawa H., Watanabe J. 2007. Observations of
Fragments Split from Nucleus B of Comet 73P/Schwassmann-Wachmann 3 with
Subaru Telescope. Publications of the Astronomical Society of Japan 59, 2, 381-386.
26
Haynes, D. R., Tro, N. J., George, S. M. 1992. Condensation and evaporation of H_2O
on ice surfaces. J. Phys. Chem. 96, 8502-8509.
Ishiguro, M., Sarugaku, Y. , Ueno, M., Miura, N., Usui, F., Chun, M.-Y., Kwon, S. M.,
2007. Dark red debris from three short-period comets: 2P/Encke, 22P/Kopff, and
65P/Gunn. Icarus 189, 1, 169-183.
Ishiguro, M. 2008. Cometary dust trail associated with Rosetta mission target:
67P/Churyumov–Gerasimenko. Icarus 193, 1, 96-104.
Jewitt, D. 2006. Comet D/1819 W1 (Blanpain): Not Dead Yet. Astron. J. 131,
2327-2331.
Kimura, H., Okamoto, H., Mukai, T. 2002. Radiation Pressure and
Poynting–Robertson Effect for Fluffy Dust Particles, Icarus, 157, 2, 349-361
Kimura, H., Mann, I., Jessberger, E. K. 2003. Composition, Structure, and Size
Distribution of Dust in the Local Interstellar Cloud. Astrophys. J. 583, 1, 314-321.
Kobayashi, H., Kawakita, H., Mumma, M. J., Bonev, B. P., Watanabe, J., Fuse, T.,
2007. Organic Volatiles in Comet 73P-B/Schwassmann-Wachmann 3 Observed during
Its Outburst: A Clue to the Formation Region of the Jupiter-Family Comets. Astrophys.
J. 668, 1, L75-L78.
Lamy, P. L., Toth, I., Weaver, H. A., 1998. Hubble Space Telescope observations of the
nucleus and inner coma of comet 19P/1904 Y2 (Borrelly). Astron. Astrophys. 337,
945-954.
Lamy, P. L., Toth, I., Fernandez, Y. R., Weaver, H. A. 2004. The sizes, shapes, albedos,
and colors of cometary nuclei. In Comets II (M.C. Festou et al., eds), University of
Arizona, Tucson, 223-264.
Landolt, A.U. 1992. UBVRI photometric standard stars in the magnitude range
the
27
11.5-16.0 around the celestial equator. Astrophys. J. 104, 340-371.
Marsden, B. G., Sekanina, Z., & Yeomans, D.K., 1973, Comets and nongravitational
forces. V, Astron. J. 78, 211-225.
Meech, K. J., Hainaut, O. R., Marsden, B. G. 2004. Comet nucleus size distributions
from HST and Keck telescopes. Icarus 170, 2, 463-491.
Miyazaki, S., Komiyama, Y., Sekiguchi, M., Okamura, S., Doi, M., Furusawa, H.,
Hamabe, M., Imi, K., Kimura, M., Nakata, F., Okada, N., Ouchi, M., Shimasaku, K.,
Yagi, M., Yasuda, N. 2002. Publications of the Astronomical Society of Japan 54, 6,
833-853.
Mukai, T., Yamamoto, T. 1982. Solar wind pressure on interplanetary dust. Astron.
Astrophys. 107, 1, 97-100.
Mukai, T. 1986, Analysis of a dirty water-ice model for cometary dust. Astron.
Astrophys. 164, 2, 397-407.
Mukai, T., Ishimoto, H., Kozasa, T., Blum, J., Greenberg, J. M. 1992. Radiation
pressure forces of fluffy porous grains. Astron. Astrophys. 262, 1, 315-320.
Prialnik, D., Bar-Nun, A. 1988. The formation of a permanent dust mantle and its effect
on cometary activity. Icarus 74, 272-283.
Prialnik, D., Sarid, G., Rosenberg, E. D., Merk, R. 2008. Thermal and Chemical
Evolution of Comet Nuclei and Kuiper Belt Objects. Space Science Reviews 138, 1-4,
147-164.
Rabinowitz, D. L. 1998, NOTE: Size and Orbit Dependent Trends in the Reflectance
Colors of Earth-Approaching Asteroids Icarus 134, 2, 342-346.
Rosenberg, E. D.; Prialnik, D. 2007. A fully 3-dimensional thermal model of a comet
28
nucleus. New Astronomy 12, 7, 523-532.
Reach, W. T., Lisse, C. M., Kelley, M. S., Vaubaillon, J. 2007. Rocket Effect For
Fragments And Meteoroids Of The Split Comet 73P/Schwassmann-Wachmann 3.
American Astronomical Society, DPS meeting #39, #54.03
Sarugaku, Y., Ishiguro, M., Pyo, J., Miura, N., Nakada, Y., Usui, F., Ueno, M. 2007.
Detection of a Long-Extended Dust Trail Associated with Short-Period Comet 4P/Faye
in 2006 Return. Publications of the Astronomical Society of Japan 59, 4, L25-L28.
Sekanina, Z. 1996, Comet 73P/Schwassmann-Wachmann 3, IAU Circ., 6301, 1
Sekanina, Z., 2005. Comet 73P/Schwassmann-Wachmann: Nucleus Fragmentation, Its
Light-Curve Signature, and Close Approach to Earth in 2006. International Comet
Quarterly, Vol. 27, pp. 225-240
Sekanina, Z., 2007. Earth's 2006 encounter with comet 73P/Schwassmann-Wachmann:
Products of nucleus fragmentation seen in closeup. In: Proceedings IAU Symposium
236. Milani, A., Valsecchi, G.B. and Vokrouhlický, D. (eds.). Cambridge: Cambridge
University Press, pp.211-220
Sosa, A., Fernandez, J. A. 2008. Cometary masses derived from non-gravitational
forces. Monthly Notices of the Royal Astronomical Society (in press).
Tancredi, G., Fernández, J. A., Rickman, H., Licandro, J. 2006. Nuclear magnitudes and
the size distribution of Jupiter family comets. Icarus 182, 2, 527-549.
Toth, I., Lamy, P. L., Weaver, H. A., 2003. Hubble Space Telescope Observations of
the Nucleus Fragment 73P/Schwassmann-Wachmann 3-B. Bulletin of the American
Astronomical Society, 35, p985
Vaubaillon, J. J.; Reach, W. T. 2006. Infrared Observations Of Comet 73p With The
Spitzer Space Telescope. Bulletin of the American Astronomical Society, Vol. 38,
29
p.490.
Villanueva, G. L., Bonev, B. P. , Mumma, M. J., Magee-Sauer, K., DiSanti, M. A.,
Salyk, C., Blake, G. A. 2006. The Volatile Composition of the Split Ecliptic comet
73P/Schwassmann-Wachmann 3: A Comparison of Fragments C and B. Astrophys. J.
650, 1, L87-L90.
Weaver, H. A., Sekanina, Z., Toth, I., Delahodde, C. E., Hainaut, O. R., Lamy, P. L.,
Bauer, J. M., A'Hearn, M. F., Arpigny, C., Combi, M. R., Davies, J. K., Feldman, P. D.,
Festou, M. C., Hook, R., Jorda, L., Keesey, M. S. W., Lisse, C. M., Marsden, B. G.,
Meech, K. J., Tozzi, G. P., West, R. 2001. HST and VLT Investigations of the
Fragments of Comet C/1999 S4 (LINEAR). Science 292, 5520, 1329-1334.
Weissman, P. R., Lowry, S. C. 2003. The Size Distribution of Jupiter-Family Cometary
Nuclei. Proc. of 34th Annual Lunar and Planetary Science Conference, abstract no.2003
Yeomans, D. K. et al. 2005. in Comets II, ed. M.C. Festou et al.(University of Arizona,
Tucson) 137
30
Table 1. Subaru/Supreme-Cam data we used in this paper
UT
(2006 May 03)
14:13
14:17-14:28
14:32
14:41-14:53
Filter Exposure Time
(sec.)
30
120
10
120
R
R
R
V
Number of
exposure
1
4
1
4
31
Figure Caption
Fig. 1. R-band composite image of 73P/Schwassmann-Wachmann 3B. These images
are oriented in the standard fashion; that is, north is up and east is to the left. The
nucleus of fragment B is masked (white area) due to the saturation. In (a), two arrows
labeled N and W denote north and west; the two arrows labeled v and denote the
directions of the comet’s orbital motion and anti-solar direction. The brightness scale
limits and contrast were modulated to emphasize the faint fragments. (b) and (d) are
enlarged views of (a) enclosed by the rectangles. (c) is obtained from a short exposure
(10 s) image. All images were processed by subtracting the 23-pixel × 23-pixel (4.6” ×
4.6”) running median images. A disconnection (see Section 2.2) is indicated by arrows.
Fig. 2. Comparison between V-band magnitude and R-band magnitude. The average
color index V-R = 0.55 is slightly redder than that of the Sun (V-R = 0.367).
Fig. 3. Observed position of the mini-fragments relative to the position of fragment B.
The dashed line denotes fragment B’s projected orbit.
Fig. 4. Histogram of mini-fragment position angle. The mean position angle is 237.5°.
The dashed line marks the Gaussian fit. The positions of fragment B’s orbit are
indicated by arrows (L denotes the leading direction and T the trailing direction).
Fig. 5. R-band magnitude (MR) vs. distance between the mini-fragments and B. The
solid line is the fitted line, MR = 1.97 log(distance) + 14.53.
Fig. 6. Comparison of the locus of synchrones (a) and syndynes (b) for dust particles
(solid lines) and mini-comets (dashed lines). (a) The synchrones are characterized by
ejection times on 16 September (rh = 3.0 AU) and 22 November (rh = 2.5 AU) 2005,
and 19 January (rh = 2.0 AU), 14 March (rh = 1.5 AU), 26 March (rh = 1.39 AU), 15
April (rh = 1.2 AU), and 24 April (rh = 1.14 AU) 2006, in a clockwise direction. (b) The
syndynes are characterized by the parameter β = 1 × 10–4, 1 × 10–3, 1 × 10–2, 1 × 10–1, in
a clockwise direction.
32
Fig. 7. Comparison of the locus of the mini-fragments with modified synchrones and
syndynes. (a) Thin dashed lines are modified synchrones characterized by ejection times
on 16 September (rh = 3.0 AU) and 22 November (rh = 2.5 AU) 2005 and 19 January (rh
= 2.0 AU), 14 March (rh = 1.5 AU), 26 March (best-fit synchrone curve of the observed
position of the mini-fragments), and 15 April (rh = 1.2 AU) 2006, rotating clockwise
from the closest curve to the dash-dotted line, which denotes the projected orbit of B.
(b) The solid lines are modified syndynes characterized by β = 1 × 10–4, 1 × 10–3, 1 ×
10–2, 1 × 10–1, rotating clockwise from the closest curve to the dash-dotted line
(projected orbit of B).
Fig. 8. Optical light curve of fragment B during the 2006 apparition. Plus signs denote
the magnitude normalized to the observer distance Δ = 1 AU. O/B and SU are the time
of the outburst expected through synchrone analysis (26 March) and the time of the
Subaru/Supreme-Cam observations (3 May). 1, 2, and 3 denote the times of the three
outbursts described by Sekanina (2007). The solid line was obtained by fitting the data
between –150 days and –65 days before the perihelion passage, HΔ = 12.7 + 15.0
log10(rh). The original
light curve data were obtained
from S. Yoshida
(http://www.aerith.net/).
Fig. 9. Radial surface brightness profile of three fragments and a reference star from
R-band data. Mini-comet 1 is apparently the most active, whereas mini-comet 3 is
inactive. The other fragment shows an intermediate brightness profile. Three thin solid
lines with slopes of –1, –2, and –3, as marked, have been included for reference.
Fig. 10. The cumulative size distribution of the mini-comets. Filled circles denote the
radius determined using standard assumptions for the geometric albedo and the linear
phase coefficient, i.e., pR = 0.04 and b = 0.04 mag deg–1. The dashed line shows the
power-law index of the differential size distribution q = –3.34, which is the result of
fitting between 12 m and 37 m.
–1 vs. radius Rc. The solid line is the result of a running median smoothing
Fig. 11. f ρc
within a window of seven nearby samples. The right axis marks the fractional active
area when ρc = 450 kg m–3.
33
Fig. 12. Cut profiles perpendicular to the comet orbit. North is to the left and south is to
the right.
34
Figure 1 (Ishiguro et al.)
35
Figure 2 (Ishiguro et al.)
36
Figure 3 (Ishiguro et al.)
37
Figure 4 (Ishiguro et al.)
38
Figure 5 (Ishiguro et al.)
39
Figure 6a (Ishiguro et al.)
Figure 6b (Ishiguro et al.)
40
Figure 7a (Ishiguro et al.)
Figure 7b (Ishiguro et al.)
41
Figure 8 (Ishiguro et al.)
42
Figure 9 (Ishiguro et al.)
43
Figure 10 (Ishiguro et al.)
44
Figure 11(Ishiguro et al.)
45
Figure 12 (Ishiguro et al.)
46
|
1112.5432 | 1 | 1112 | 2011-12-22T20:04:43 | Outward migration of a super-Earth in a disc with outward propagating density waves excited by a giant planet | [
"astro-ph.EP"
] | In this paper we consider a new mechanism for stopping the inward migration of a low-mass planet embedded in a gaseous protoplanetary disc. It operates when a low-mass planet (for example a super-Earth), encounters outgoing density waves excited by another source in the disc. This source could be a gas giant in an orbit interior to that of the low-mass planet. As the super-Earth passes through the wave field, angular momentum is transferred to the disc material and then communicated to the planet through coorbital dynamics, with the consequence that its inward migration can be halted or even reversed. We illustrate how the mechanism we consider works in a variety of different physical conditions employing global two-dimensional hydrodynamical calculations. We confirm our results by performing local shearing box simulations in which the super-Earth interacts with density waves excited by an independent harmonically varying potential. Finally, we discuss the constraints arising from the process considered here, on formation scenarios for systems containing a giant planet and lower mass planet in an outer orbit with a 2:1 commensurability such as GJ876. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 24 (2010)
Printed 11 June 2021
(MN LATEX style file v2.2)
Outward migration of a super-Earth in a disc with outward
propagating density waves excited by a giant planet
E. Podlewska-Gaca1,3(cid:63), J. C. B Papaloizou2† and E. Szuszkiewicz1,3‡
1Institute of Physics and CASA*, University of Szczecin, ul. Wielkopolska 15, 70-451 Szczecin, Poland
2Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Wilberforce Road,
Cambridge CB3 0WA, United Kingdom
3Kavli Institute for Theoretical Physics, University of California, Santa Barbara, California 93106, USA
Accepted; Received; in original form
ABSTRACT
In this paper we consider a new mechanism for stopping the inward migration of a
low-mass planet embedded in a gaseous protoplanetary disc. It operates when a low-
mass planet (for example a super-Earth), encounters outgoing density waves excited
by another source in the disc. This source could be a gas giant in an orbit interior to
that of the low-mass planet. As the super-Earth passes through the wave field, angular
momentum is transferred to the disc material and then communicated to the planet
through coorbital dynamics, with the consequence that its inward migration can be
halted or even reversed.
We illustrate how the mechanism we consider works in a variety of different phys-
ical conditions employing global two-dimensional hydrodynamical calculations. We
confirm our results by performing local shearing box simulations in which the super-
Earth interacts with density waves excited by an independent harmonically varying
potential. Finally, we discuss the constraints arising from the process considered here,
on formation scenarios for systems containing a giant planet and lower mass planet in
an outer orbit with a 2:1 commensurability such as GJ876.
Key words: planets and satellites: formation - methods: numerical, planet-disc in-
teractions
1 INTRODUCTION
The orbital migration of low-mass planets in protoplanetary
discs may play an important role in shaping planetary sys-
tems. In our previous studies (Papaloizou & Szuszkiewicz
2005; Podlewska & Szuszkiewicz 2008, 2009; Papaloizou &
Szuszkiewicz 2010) we have concentrated on the possibil-
ity of the occurrence of a resonant structure during early
phases of the evolution of the planetary system. We have
considered two-planet systems, containing at least one low-
mass planet of several Earth masses, embedded in a gaseous
disc. In Podlewska & Szuszkiewicz (2008) we found that in a
system with one low-mass planet (described here as a super-
Earth) and a gas giant, the formation of commensurabilities
readily occurs when the migration of the two planets is con-
vergent with the gas giant being on the larger orbit. How-
ever, as shown in Podlewska & Szuszkiewicz (2009), this is
no longer true if the migration is convergent but the gas gi-
ant is on the smaller orbit. In this case the super-Earth never
(cid:63) E-mail: [email protected] (EP)
† E-mail:[email protected] (JP)
‡ E-mail: [email protected] (ES)
c(cid:13) 2010 RAS
gets close enough to the location of the 2:1 commensurabil-
ity in order to be caught in it. The super-Earth is initially
found to migrate towards the central star but at some point
stops or even starts to migrate slowly outward. However,
no explanation of this phenomenon was given. The aim of
our paper is to explore this behavior more fully using both
global and local shearing box simulations with the aim of
delineating the mechanism responsible for the outward mi-
gration, showing how it depends on model parameters such
as the mass of the giant planet and the disc aspect ratio.
Since it was realized that the fast migration of low-
mass planets (Ward 1997) could pose a serious challenge
to scenarios for planet formation, possibilities for slowing
down such migration have been the subject of intensive stud-
ies. A number of potential mechanisms have been found by
focusing on a single low-mass planet embedded in a pro-
toplanetary disc. These mechanisms involve, among oth-
ers, entry into a magnetospheric cavity close to the star
(Lin et al. 1996), effects arising from the orbital eccentric-
ity of a protoplanet (Papaloizou & Larwood 2000), effects
due to the possible eccentricity of the protoplanetary disc
(Papaloizou 2002), magnetic fields (Terquem 2003), MHD
2
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
turbulence (Laughlin, Steinacker & Adams 2004; Nelson
& Papaloizou 2004; Johnson, Goodman & Menou 2006;
Adams & Bloch 2009), sudden jumps in disc state vari-
ables (Menou & Goodman 2004; Matsumura, Pudritz &
Thommes 2006), corotation torques (Masset et al. 2006;
Paardekooper & Papaloizou 2009a,b), disc thermodynamics
(Paardekooper & Mellema 2006; Baruteau & Masset 2008a;
Kley & Crida 2008; Paardekooper & Papaloizou 2008; Kley,
Bitsch & Klahr 2009; Paardekooper et al. 2010; Hasegawa
2010; Paardekooper et al. 2011; Yamada & Inaba 2011) and
instabilities of partial gap edges in low viscosity discs (Li et
al. 2009; Yu et al. 2010).
However, there may be mechanisms for affecting migra-
tion that depend on having a system of planets. For exam-
ple, Ogihara et al. (2010) have suggested the possibility of
"an eccentricity trap" operating for a resonantly interacting
convoy of planets in a way to halt type I migration near the
inner edge of a protoplanetary disc. Thommes (2005) sug-
gested that gas giants may be efficient at capturing low-mass
planets in their exterior mean-motion resonances, forming in
this way a barrier in a protoplanetary disc "a safety net for
fast migrators". A follow up study based on numerical sim-
ulations of the disc planet interactions has been been pre-
sented by Pierens & Nelson (2008). They found that planets
with masses in the range 3.5-20 M⊕ become trapped at the
edge of the gap formed by the giant planet, while more mas-
sive planets are captured into resonance.
They noted that the positive surface density gradient re-
sembled the configuration discussed by Masset et al. (2006)
for their planet trap. Accordingly they raised the possibility
that corotation torques operating as in a tidally undisturbed
disc could be responsible. However, the configuration differs
significantly from that of Masset et al. (2006). The surface
density profile in the latter's planet trap is maintained by
a viscosity that increases rapidly inwards. In the absence of
a viscous angular momentum flux produced by the action
of viscous stresses at an inner boundary, acceleration of the
accretion flow would occur, that in turn causes a decrease
in the surface density in a steady state. However, in a disc
tidally disturbed by a giant planet, the profile is maintained
by the outward tidal transport of angular momentum pro-
duced by dissipating density waves rather than by either
an accretion flow, or effects of comparable magnitude re-
sulting from applied viscous stresses. The wave transport
also occurs in the coorbital region of a migrating planet and
thus coorbital torques may be expected to behave differently
from those occurring in an undisturbed disc. Notably, having
an external source of angular momentum for the coorbital
region means that positive corotation torques can be sus-
tained without either the action of applied viscous stresses,
or the accretion of exterior fresh material into the coorbital
zone followed by its loss into the inner regions, which would
otherwise be required. Put another way, an external source
of angular momentum for the coorbital zone could prevent
corotation torques from saturating as would be expected for
a tidally undisturbed disc with small applied viscosity. Thus,
as seen in our simulations, long term positive torques may
be communicated to the planet even when the nominal vis-
cosity is set to zero. In this way, even though there are some
similarities, effects seen in a disc that is tidally disturbed by
a giant planet may differ from those found in a disc with a
surface density profile maintained through the application
of a variable viscosity.
The mechanism, we describe in this paper, operates in a
similar physical situation where a low mass planet migrates
in a disc which contains trailing density waves excited by
another planet. A specific configuration of this type that
we consider has a gas giant in an interior orbit with a low
mass planet in an exterior orbit around a central mass (see
Podlewska & Szuszkiewicz 2009). However, local simulations
that we perform indicate that the essential features are a sin-
gle planet immersed in a field of propagating density waves
produced by an external source. The waves are associated
with an angular momentum flux which can enter the coor-
bital region of the low mass planet and transfer angular mo-
mentum to the disc material there through the dissipation
of shocks. This in turn can transfer angular momentum to
the low mass planet through coorbital dynamics. It is im-
portant to note that this does not depend on vortensity gra-
dients as do standard corotation torques (e.g. Masset et al.
2006; Paardekooper & Papaloizou 2009a,b). In addition be-
cause the waves act as a source of angular momentum from
outside the coorbital zone and the transfer involves viscos-
ity independent shock wave dissipation, the usual satura-
tion considerations relevant to standard corotation torques
do not apply.
This paper is organized as follows. In Section 2 we de-
scribe the physical model and give the basic governing equa-
tions. Then in Section 3 we give a brief review of the type I
migration theory applicable to an isolated super-Earth em-
bedded in a protoplanetary disc without propagating den-
sity waves. We go on to present simulations of a super-Earth
embedded in a protoplanetary disc in which density waves
excited by an interior giant planet propagate in Section 4.
We show that in the presence of outward propagating trail-
ing density waves, the inward type I migration of a super-
Earth can be reversed with the consequence that attain-
ment of a 2:1 commensurability by migrating inwards from
large radii is prevented. We go on to describe local shearing
box simulations of a planet embedded in a disc with density
waves excited by an imposed harmonically varying potential
in Section 5. We find similar results to those for the global
simulations. The waves are found to induce angular momen-
tum transport in the disc material that is then passed on to
the planet. In Section 6 we consider this mechanism, which
leads to the outward migration of the super-Earth in the
global simulations, from the point of view of the conserva-
tion of energy and angular momentum. We then consider
some consequences of the wave-planet interaction for the
resonance capture into an outer 2:1 commensurability of a
super-Earth by a gas giant in Section 7. In general density
waves excited by the giant planet tend to prevent this, in-
dicating that such commensurabilities , as seen for example
in GJ876, may have formed through planetesimal migration
after the gas disc dispersed. Finally in Section 8 we give our
conclusions.
2 BASIC EQUATIONS AND MODEL
We consider two planets of masses M1 and M2 orbiting a
central star of mass M∗. We adopt a cylindrical coordinate
system (r, φ, z) where z is the vertical coordinate increas-
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
3
Figure 1. The left panel shows the evolution of the semi-major axis of a super-Earth in a presence of a gas giant with
one Jupiter mass (black curve) and two Jupiter masses (grey curve). The right panel shows the variation of the torque
acting on the super-Earth: coming from the disc matter located in the vicinity of the planet, in the region between r = 1.4
and r = 1.7. The timescale covers many periods of the super-Earth, which is located at r ∼ 1.545 which is close to the
azimuthally averaged surface density maximum that occurs slightly exterior to the gap edge. Note during this time period,
the giant planet is located at r ∼ 0.9, to which it has migrated.
Figure 2. (Left panel) The semi-major axis of the super-Earth (top) and its distance from the central star (bottom) as a function of
time. The time interval in each picture corresponds roughly to one orbital period of the super-Earth. (Right panel) The torque acting
on the super-Earth: total torque (top), the one coming from the disc matter located in the close vicinity of the planet, namely between
r = 1.4 and r = 1.7 (bottom). The vertical lines in the right bottom panel indicate particular moments of time for which the planet
positions in the disc are shown in Fig. 3.
ing in the direction normal to the disc plane for which the
unit vector is k. We integrate vertically to obtain a two-
dimensional flat disc model, for which the governing hydro-
dynamic equations in an inertial frame are the continuity
equation
= −∇ · (Σu),
∂Σ
∂t
and the equation of motion
∂u
∂t
+ u · ∇u = − 1
Σ
∇P − ∇Φ + fν ,
(1)
(2)
where the vertically integrated pressure is P = c2
s(r)Σ with
cs(r) = h(GM∗/r)1/2, and h being the assumed constant
j=1
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
.
+
GMjr
r2
j
cos (φ − φj)
(3)
disc aspect ratio. This is related to the putative disc semi-
thickness, H(r), through h = H(r)/r. Thus a locally isother-
mal equation of state with sound speed cs(r) is adopted. The
surface density is Σ and u is the velocity. The viscous force
per unit mass is fν . The detailed form of this is given in
Nelson et al. (2000). The effective gravitational potential is
given by
Φ = − GM∗
r
−
2(cid:88)
(cid:113)
GMj
r2 + r2
j − 2rrj cos(φ − φj) + 2
p
1.45 1.5 1.55 1.6 1.65 1.7 1.75 0 1000 2000 3000 4000 5000 6000 7000 8000 9000semi-major axistime-0.0001-5e-05 0 5e-05 0.0001 0.00015 4900 4950 5000 5050 5100 5150torque from r=1.4 to r=1.7time 1.536 1.538 1.54 1.542 1.544 1.546 1.548 1.55 1.552 4904 4906 4908 4910 4912 4914 4916 4918 4920semi-major axistime-0.002-0.0015-0.001-0.0005 0 0.0005 0.001 0.0015 0.002 0.0025 4904 4906 4908 4910 4912 4914 4916 4918 4920total torquetime 1.53 1.535 1.54 1.545 1.55 1.555 1.56 4904 4906 4908 4910 4912 4914 4916 4918 4920rtime-6e-05-4e-05-2e-05 0 2e-05 4e-05 6e-05 8e-05 0.0001 0.00012 4904 4906 4908 4910 4912 4914 4916 4918 4920torque from r=1.4 to r=1.7time12345674
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
The last term on the right hand side is the indirect term
accounting for the acceleration of the primary. In the above,
the cylindrical coordinates of the planets that are assumed
to be confined to the disc plane are (ri, φi, 0), i = 1, 2 and
the softening length is p. Note that a non zero value of p
can be thought of as accounting very approximately for the
effects of vertical structure. For the simulations presented
here, we adopted p = 0.024r. Thus for h = 0.03 the soften-
ing length is 0.8H. The corresponding quantities for h = 0.05
and h = 0.07, for which we have also performed simulations
are 0.48H and 0.34H respectively.
Setting the Keplerian angular velocity to be, Ω(r) =
(GM∗/r3)1/2, the unit of time is taken to be Ω(r1,0)−1. Here
the gravitational constant is G, M∗ denotes the mass of the
star and r1,0 denotes the initial radial position of the giant
planet. The unit of time corresponds to (1/2π) times the
orbital period of the initial orbit of the inner giant planet.
The adopted unit of length is the initial orbital radius of the
giant planet which can be assumed to correspond to 5.2 au.
However, it may be scaled from this to arbitrary values by
applying a scaling factor, λ, provided that the disc surface
density is scaled by λ−2 preserving the mass scale. The unit
of time then scales as λ3/2.
We consider discs which undergo near-Keplerian rota-
tion with H being small compared to r. We have adopted two
initial surface density profiles for the disc. For the standard
initial disc, which was adopted below unless stated other-
wise, the profile was taken to be uniform with Σ = Σ0 in the
vicinity of the super-Earth with tapers at inner and outer
radii prescribed according to
Σ = (0.1 + 4.5(r − rmin))Σ0
Σ = Σ0
−1.5
rmin + 0.2 (cid:54) r < 1.65,
r < rmin + 0.2,
for
for
Σ = Σ0 × (r/1.65)
r > 1.65,
for
(4)
where rmin being the inner edge of the disc was taken to be
0.33.
Some of our simulations were carried out for an initial
disc with a more extensive uniform surface density region,
with as before, exterior and interior tapers but also a region
of low surface density in the vicinity of the orbital radius of
the giant planet, corresponding to a gap. The surface density
was given by
Σ = (0.1 + 4.5(r − rmin))Σ0
Σ = Σ0
−11.25
rmin + 0.2 (cid:54) r < 0.7,
r < rmin + 0.2,
for
for
Σ = Σ0 × (r/0.65)
for
Σ = 0.001Σ0
Σ = Σ0 × (r/1.43)11.25
Σ = Σ0
−1.5
Σ = Σ0 × (r/4.3)
for
for
for
for
0.7 (cid:54) r < 0.75,
0.75 (cid:54) r < 1.15,
1.15 (cid:54) r < 1.43,
1.43 (cid:54) r < 4.3,
r > 4.3.
(5)
This disc is referred to below as the one having a more ex-
tensive uniform surface density region. We remark that, due
to nonlinear perturbations from the giant planet the sur-
face density profile, in the vicinity of and interior to the low
mass planet, tends to relax quickly to a form that does not
depend on the initial profile used. However, differences at
larger radii may remain throughout the simulations.
In practice we adopted Σ0 = 6 × 10−4M(cid:12)/(5.2au)2 which,
for M∗ = M(cid:12), corresponds to the minimum mass solar neb-
ula (MMSN) with two Jupiter masses contained within a
circular area of radius equal to 5.2au. Note that results
may be scaled to different values of M∗ by multiplying the
planet masses by M∗/M(cid:12) and surface densities and radii by
(M∗/M(cid:12))1/3. The mass M1 is taken to be that of a giant
planet and the mass M2 is taken to be that of a super-Earth.
All planets are initialized on circular orbits.
We used the Eulerian hydrodynamic code NIRVANA
(Ziegler 1998) to solve the governing equations as was done
in our previous papers (Papaloizou & Szuszkiewicz 2005;
Podlewska & Szuszkiewicz 2008, 2009). The details of the
numerical scheme can be found in Nelson et al. (2000). The
computational domain in the radial direction extends from
rmin = 0.33 to rmax, which has been taken to be in the range
4 − 5. The azimuthal angle ϕ lies in the interval [0, 2π]. The
disc domain is partitioned into grid cells in such a way that
a single grid cell is approximately square with sides parallel
to the radial and azimuthal directions being approximately
equal to 0.01 in dimensionless units. The radial boundaries
were taken to be open.
When calculating the gravitational forces on the plan-
ets due to the disc, we include the matter contained in the
planet's Hill spheres. We remark that in practice only the
super-Earth is enveloped by the disc and for such a low-
mass planet, the influence of the material inside Hill sphere
can in any case be neglected. The self-gravity of the disc is
neglected.
We have run our numerical experiments with disc as-
pect ratio, h, ranging from 0.03 to 0.07 and constant kine-
matic viscosity ν. Unless stated otherwise, we have adopted
ν = 2× 10−6 in dimensionless units (this corresponds to the
standard α parameter being equal to 2.2×10−3 for h = 0.03,
8 × 10−4 for h = 0.05 and 4.1 × 10−4 for h = 0.07).
3 AN ISOLATED SUPER-EARTH EMBEDDED
IN A PROTOPLANETARY DISC
The migration of low-mass planets in a gaseous disc has been
calculated from the linear response of the disc to the pertur-
bation caused by the presence of a planet (e.g. Tanaka et al.
2002). This procedure is reasonable if coorbital torques do
not play a role. The latter depend on the disc surface density
and temperature profiles and thermodynamics. When they
are important, non-linear effects play a role (Paardekooper
et al. 2011; Yamada & Inaba 2011). According to the recent
studies, a single low-mass planet can migrate with a whole
range of speeds, both inwards and outwards, depending on
the assumed physical and structural properties of the disc
in which it is embedded.
For the disc models we consider, a super-Earth with the
mass of 5.5 M⊕ orbiting the Solar-mass star embedded in
such a disc undergoes inward migration with a speed similar
to that derived from the linear treatment of Tanaka et al.
(2002). The situation is very different if the super-Earth
migrates in a disc in which density waves excited by a gas
giant are present and have a significant interaction with it.
We go on to consider this situation below.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
5
Figure 3. The surface density contours in the disc at a time denoted by 1. The wake associated with the super-Earth can
be clearly seen. The planet positions at six additional times denoted 2 − 7 are correspondingly marked to indicate their
relation to the wave pattern excited by the Jupiter mass planet. The times correspond to those indicated in Fig. 2 (right
bottom panel), for which the values of torques have been given. The Jupiter mass planet is at fixed azimuthal position as
would be situation in a frame that corotates with it.
4 A SUPER-EARTH EMBEDDED IN A
PROTOPLANETARY DISC IN WHICH
DENSITY WAVES EXCITED BY AN
INTERIOR GIANT PLANET PROPAGATE
Before considering the migration of the super-Earth, we
briefly review some well known relevant properties of the
density waves excited in a gaseous disc by an orbiting giant
planet.
4.1 Density waves excited by a giant planet
It is well known that an orbiting planet or satellite excites
density waves in a nearby gaseous disc that is orbiting the
same central body (Goldreich & Tremaine 1979). The den-
sity waves are launched at Lindblad resonances and prop-
agate away from the planet where they eventually shock
and dissipate. Because of this they take the form of trailing
waves propagating away from the planet which are always
associated with outward angular momentum transport (e.g.
Lynden-Bell & Kalnajs 1972; Balbus & Papaloizou 1999).
For a non migrating or slowly migrating protoplanet
embedded in a disc without density waves viewed in a frame
rotating with the mean angular velocity of the planet, there
are closed streamlines that represent a coorbital flow. Each
of these can be thought of as constituting a single coorbital
cell that occupies almost the full 2π in azimuth, with the
protoplanet located in the excluded region (see for example
the 'librating streamlines' in Fig. 5 of Masset & Papaloizou
(2003)). This of course is the same situation that occurs
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
in particle dynamics without pressure where the librating
orbits correspond to horseshoe orbits.
When density waves are present, the coorbital flows may
be significantly affected by general pressure forces and shock
waves such they depart from the normal horseshoe form.
However, if fluid elements still remain trapped in the coor-
bital region, dissipative effects acting on the coorbital flow
can lead to a net force or torque acting on the super-Earth,
a phenomenon that will be discussed further in Sections 4.2
and 6.2 below.
This is analogous to the situation that occurs for parti-
cles that are subject to secular dissipative forces that cause
them to slowly migrate in the absence of the protoplanet.
When a non migrating low mass protoplanet that induces
horseshoe trajectories is present, particles on those trajec-
tories can become trapped, communicating the dissipative
forces/torques that act on them to the protoplanet (e.g. Der-
mott et al. 1980).
In order for this phenomenon to result in a positive
torque on the protoplanet, the local dissipation of the waves,
in practice mostly through shocks, should produce a posi-
tive torque on the disc material, as is indeed expected for
trailing density waves excited by an interior giant planet.
Furthermore this transport should exceed any counteract-
ing effects produced by a disc viscosity on average, in order
for a net positive torque to be communicated to the proto-
planet. Thus the torques associated with the dissipation of
the density waves should exceed viscous torques, implying
6
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 4. The surface density (left panels) and streamlines in localized disc segments (right panels). The uppermost, middle and lowermost
panels correspond to the first, second and third moments of time denoted in the lower right panel of Fig. 2 and Fig. 3.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
7
Figure 5. The surface density (left panels) and streamlines in localized disc segments (right panels). The uppermost, middle and lowermost
panels correspond to the fourth, fifth and sixth moments of time denoted in the lower right panel of Fig. 2 and Fig. 3.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
8
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 6. The surface density contours for the disc with aspect ratio h=0.03 (left), h=0.05 (middle), and h=0.07 (right) at
two different moments of the evolution, namely at about 10800 time units (three upper panels) and 45000 time units (three
middle panels). The density waves weaken at later times because the gap widens and the disk tends to expand outwards
along with the super-Earth. The evolution of the semi-major axis of a super-Earth in the disc with aspect ratio h=0.03
(black), h=0.05 (dark grey), and h=0.07 (light grey) (bottom panel).
Figure 7. The azimuthally averaged surface density as a function of radius for the disc with aspect ratio h=0.03 (left),
h=0.05 (middle), and h=0.07 (right). The position of the super-Earth is indicated by the black dots. The snapshots are taken
after 6280, 12556, 25132 and 37600 the time units. The surface density level at the planet location decreases monotonically
with time.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 0 10000 20000 30000 40000 50000semi-major axistime 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0.0008 0.0009 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5surface densityradial distance from the star 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5surface densityradial distance from the star 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0.0008 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5surface densityradial distance from the starA super-Earth migrating in a disc with density waves excited by a gas giant
9
that in a balanced state, tidal effects exceed viscous effects
with the balance communicated to the super-Earth.
4.2 Interaction of an embedded super-Earth with
propagating density waves
It was shown in Podlewska & Szuszkiewicz (2009) that a
super-Earth in an orbit exterior to the giant, migrating
in the disc towards a Jupiter-mass planet, never got close
enough to the giant in order to be caught in a 2:1 com-
mensurability. Before it reached this location, it began to
migrate outwards. Let us examine the stage of the evolution
when the super-Earth migrates outwards.
We consider a disc with h = 0.03, ν = 2 × 10−6 and
Σ0 = 0.0006 in dimensionless units. The giant planet was
taken to be of one Jupiter mass and the initial orbital ra-
dius of the super-Earth was taken to be 1.62. Tests have
shown that when the super-Earth is located beyond the 2:1
resonance with the interior giant its migration behaviour is
not very much affected by the mutual gravitational interac-
tion of the planets or its initial orbital radius. Accordingly,
to clarify that it is the presence of the density waves that
induces outward migration, we perform studies in which this
interaction is switched off. The interaction of both planets
with the central star and the disc remains intact.
The evolution of the semi-major axis of the super-Earth
is shown as a function of time in Fig. 1. This shows that af-
ter an initial phase in which the semi-major axis decreases
on average it eventually increases on average, the transition
occurring around t = 4000. At this point we emphasize that
this ultimate outward migration is not a consequence of the
particular disc model or initial conditions adopted (see be-
low).
In order to see how the average torque acting on the
super-Earth can then be positive, we consider the behaviour
of the semi-major axis of the super-Earth and its distance
from the central star. These are plotted for a time span of
one typical orbital period of the super-Earth, starting at
t ∼ 4906, where the mean migration rate is outwards, in
Fig. 2. In Fig. 2 we also show the torques acting on the
super-Earth arising from both the whole disc and only the
region located in the domain 1.4 < r < 1.7 that is close to
the planet. The average positive torque exhibited over the
one orbital period of the super-Earth occurs for subsequent
orbital periods at approximately the same level (see Fig. 1)
although the torque itself shows a long period modulation
implying some unsteadiness.
We have calculated the values of the torques acting on
the planet at seven times which are indicated by the vertical
lines in Fig. 2. To illustrate the location of the super-Earth
with respect to the density waves excited in the disc by
the giant at these times, we indicate the planet positions
as seen in a frame that corotates with the giant on the top
of the surface density contours at a single moment of time
in Fig. 3. Here we remark that the density wave pattern is
approximately fixed as seen in a frame that corotates with
the giant. Note too that at this time Fig. 3 indicates that
the giant planet has migrated inwards to r ∼ 0.9.
In Figs. 4 - 5 we show the surface density contours in a
close vicinity of the planet at the first six times indicated in
the lower right panel of Fig. 2. These fall within one orbital
period of the super-Earth. We also show the streamlines
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
both in the vicinity of the planet and for some cases also
a coorbital azimuthal domain separated from the planet.
This is to confirm that the streamlines do close. For all but
the second time, the torque exerted in the vicinity of the
super-Earth is positive. Although it is sometimes difficult to
discern because of cancellation effects, this is in general seen
to be expected from the form of the density contours. For
example at the first time there is a surface density excess
to the right of the planet that contributes a positive torque,
while at the second time there is a surface density excess
to the left of the super-Earth that contributes a negative
torque.
The coorbital streamlines are closed. At the fourth time
they resemble the form associated with steady state horse-
shoe orbits with an X point near the planet's location. Thus
there are oppositely directed turns on either side of the
planet. However, because of the presence of density waves
the pattern is different in other cases, for example at the first
and the sixth times, the planet is inside the circulating re-
gion with an X point significantly shifted from the planet at
φ ∼ 1.1 and 1.5 respectively. Because the situation is a dy-
namic one, it cannot be said that fluid elements follow these
streamlines but they do indicate that they are turned close
to the planet at least in some cases. Just as in the standard
horseshoe case this is associated with angular momentum
exchange with the planet even though the situation is more
complex because of the time dependence and the deflection
of the streamlines by density waves altering their form.
We have performed a series of simulations in order to
investigate the dependence of the outward torque on the
physical set up. The evolution of the semi-major axis of
a super-Earth for giant planet masses of one Jupiter mass
(black curve) and two Jupiter masses (grey curve) is shown
in Fig. 1. In these calculations, the gas giant is allowed to
migrate and the super-Earth does not interact gravitation-
ally with a gas giant. This ensures that the torque acting on
the super-Earth arises entirely from the disc material.
We remark that density wave excitation is expected to
be enhanced by placing a more massive giant in the disc.
The initial inward migration of the super-Earth is slower in
the case with the two Jupiter mass giant. Also in this case,
outward migration causes the super-Earth to reach larger
radii at equal large times. This is consistent with the idea
that the outward torque is produced by some fraction of the
outward angular momentum transfer rate associated with
density waves excited by the giant. This torque competes
with the torques associated with standard type I migration
to determine how far the super-Earth can migrate outwards.
We have also performed calculations in which the super-
Earth is embedded in discs of different aspect ratio and for
which the giant planet was not allowed to migrate.
The pitch of the spiral arms excited by the giant planet
depends on the inverse Mach number, or equivalently the
aspect ratio h = H/r of the disc. The smaller the aspect
ratio the more tightly the spiral arms are wound. In this case
the larger Mach number results in the waves more readily
becoming nonlinear and forming shocks.
In Fig. 6 we show the disc surface density contours for
discs with aspect ratio h = 0.03 (left panel), h = 0.05 (mid-
dle panel), and h = 0.07 (right panel) for a one Jupiter
mass giant planet. The snapshots are taken after 10800 and
45000 time units. It can be seen that as the aspect ratio
10
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
increases, the density waves launched by the giant planet
become weaker and less tightly wound.
The migration of the super-Earth in these three differ-
ent environments is shown in Fig. 6 (bottom panel). The
initial orbital radius was 1.64 in each case. The black curve
shows the evolution of the semi-major axis of the super-
Earth embedded in the disc with aspect ratio 0.03, the dark
grey curve in the disc with h = 0.05, and the light grey curve
in the disc with aspect ratio 0.07.
For the disc with h = 0.03 the interaction of the nonlin-
ear density waves with the super-Earth causes it to increase
its radius from r = 1.6 to r = 1.72 in approximately 4000
time units. For the disc with h = 0.05 the density waves are
weaker and interact with the planet less strongly than in the
previous case. Indeed, the outward migration is not so fast
and proceeds almost at the same rate during the whole evo-
lution. For the h = 0.07 the density waves are even weaker
as is their interaction with the planet with the result that it
migrates outward very slowly.
To complete our discussion of these simulations we
present the azimuthally averaged surface density at different
times during the dynamical evolution of the super-Earth for
the discs with the different aspect ratios. They are shown in
Fig. 7 after 6280, 12556, 25132 and 37600 time units. The
azimuthally averaged surface density at the planet location
decreases monotonically with time. The planets continue to
migrate outwards but they remain in the neighbourhood of
the azimuthally averaged surface density maximum in all
these runs. This is an indication that the angular momen-
tum transfer due to the density waves continues to widen
the gap as well as move the planet outwards. This might be
expected if a positive torque acting on the planet requires a
non zero outward angular momentum flux carried in density
waves. The presence of such a flux might be expected to be
associated with gap production interior to the planet.
In Section 5 we study the effect of density waves on
planet migration by doing a local calculation of a planet
in the presence of forced waves. This approach has the ad-
vantage that a very deep gap does not necessarily result.
However, the forcing of the planet is similar to what we see
in discs where the waves are excited by a giant planet. Thus
we gain a strong indication in favour of the hypothesis that
the outward migration studied here is a consequence of the
direct interaction of a low-mass planet with the spiral waves.
We have also investigated how the evolution of the
super-Earth depends on its initial location in the disc. In
the left panel of Fig. 8 we show the results of simulations
for which the initial orbital radii were 1.64 and 1.9. These
show that regardless of the initial orbital radius, the radial
position of the planet always approaches the same curve.
Another important issue is the dependence of the migra-
tion behaviour of a low-mass planet on its mass. For this pur-
pose we have performed simulations with embedded planet
masses of one Earth mass, 5.5 Earth masses, and 10 Earth
masses embedded in a disc with aspect ratio h = 0.03. Their
initial orbital radii were 1.64. The results are also shown in
Fig. 8 (right panel). The giant planet, of one Jupiter mass, is
not allowed to migrate and, as previously, the planets don't
interact with each other gravitationally. The results differ
from what would be expected assuming that the outward
torque occurs as a direct result of interaction with the spi-
ral wave. In that case it would be proportional to the planet
mass. As the torques arising from disc planet interactions are
proportional to the square of the mass in the linear regime,
the planet with the lowest mass should be pushed out the
most rapidly. This is not the case in our calculations, which
indicate the opposite in the early stages.
All the numerical experiments discussed so far indicate
that low-mass planets reverse their inward migration at a lo-
cation, which is very close to a strongly peaked azimuthally
averaged surface density maximum. This feature of our cal-
culations appears due to the particular choice of the initial
standard surface density profile Eq.(4). To illustrate this,
we replace this profile by the more extensive one given by
Eq.(5). The outer edge of this disc is now far enough away
so that it does not affect the surface density distribution
in the vicinity of the planet. The initial profile incorporates
a gap so that we eliminate the surface density bumps that
appear at gap edges as a result of the gap opening process.
In Fig. 9 we show the evolution of the semi-major axis of a
planet that starts from the initial value of 1.9, in both the
standard disc and the more extensive disc given by Eq.(5).
The aspect ratio is taken to be h = 0.05 in both cases. The
azimuthally averaged surface density profiles for these two
cases are shown in Fig. 10. The snapshots are taken after
3100, 6200 and 10000 time units. At the last time there is
slow outward migration in both cases. There is a significant
difference in the behaviour of the super-Earth migration.
The super-Earth initially migrates much more slowly in the
case starting with surface density distribution given by Eq.
(5), which is in line with the behaviour of standard corota-
tion torques which are expected to be stronger for the case
with more rapidly decreasing azimuthally averaged surface
density (eg. Masset et al. 2006; Paardekooper & Papaloizou
2009a). However, it could also be a consequence of the tidal
effects of the giant being less effective at larger radii in that
case.
Let us see how the results of our previous experiments
depend on the adopted surface density profile of the disc.
In the left panel of Fig. 11 we show the evolution of the
semi-major axis of the super-Earth for gas giant masses of
one Jupiter mass and two Jupiter masses. As before, other
things being equal, the super-Earth outward migration rate
is larger if the gas giant has a larger mass.
Now, we come back to the simulations carried out for
three different masses of the low-mass planet (Fig. 8) and
redo them starting with the more extensive surface density
profile Eq.(5). We present a comparison of the evolution of
planets with masses 1, 5.5 and 10 Earth masses in a disc with
aspect ratio h = 0.05. In Fig. 11 (right panel) we show the
evolution of their semi-major axes. We see that the larger is
the mass of the planet, the greater is the extent of the initial
inward migration and the larger the mass of the planet, the
faster is the subsequent outward migration. The second of
these conclusions could be inferred from the previously dis-
cussed low-mass planet evolution in the disc with the stan-
dard initial surface density profile (Fig. 8) immediately after
the reversal of the direction of the migration.
4.3 The effect of varying viscosity
We have checked that the similar results as found above
occur for viscosity ν in the normally considered range
ν (cid:54) 10−5. For example Fig. 12 compares the migration be-
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
11
Figure 8. The left panel shows the evolution of the semi-major axis of a 5.5 Earth mass planet for which the initial distance
from the central mass was 1.64 (black line) and 1.9 (grey line). The aspect ratio of the disc is h=0.05. The right panel shows
the evolution of the semi-major axis of a low mass planet with a mass of 1 Earth mass (light grey curve), 5.5 Earth masses
(black curve) and 10 Earth masses (dark grey curve). The disc aspect ratio h = 0.03.
Figure 9. The evolution of the semi-major axis of a super-Earth in a disc with standard initial surface density distribution
(black curve) and in a disc with the more extensive initial surface density distribution (grey curve).
haviour and azimuthally averaged surface density profile at
t = 12000 for ν = 2 × 10−6 and ν = 10−5. For these simula-
tions the giant planet was one Jupiter mass and they were
initiated with the extended surface density profile.
The surface density profile is smoother in the case with
larger viscosity and as might be expected, more material
leaks interior to the giant planet orbit in this case.
We have also confirmed that similar stalled or outward
migration occurs when the applied viscosity is set to zero.
We remark that De Val-Borro et al. (2007) find that nu-
merical diffusion in NIRVANA under conditions like those
adopted here corresponds to at most ν ∼ 10−7 at a partic-
ular grid location , while usually amounting to much less.
Thus the lack of the smoothing effect provided by even a
small applied viscosity results in an initial condition depen-
dent oscillatory form for the azimuthally averaged surface
density profile beyond the gap that comes about from the
angular momentum transport induced by the tidal effect of
the giant planet planet acting on the initial disc. As there
is no long term evolution of the outer disc due to the action
of an internal viscosity, these features persist. This in turn
leads to some differences in the migrational behaviour of the
low mass planet when compared to cases with a non zero ν.
In these simulations we adopted an initial disc with the ex-
tended surface density profile and aspect ratio h = 0.05. The
mass of the super-Earth was 5.5 Earth masses.
We have considered the case of a non migrating Jupiter
mass giant planet with an initial radius for the super-Earth,
r = 1.65. The results of this simulation as well as the oth-
ers undertaken with zero applied viscosity are illustrated
in the lower panels of Fig. 12. The low mass planet mi-
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
grated inwards initially but then subsequently halted with
radius slightly exceeding 1.6 dimensionless units. This situ-
ation could be followed for about 1500 orbits of this planet.
Because the azimuthally averaged surface density in these
cases has an oscillatory structure as a function of radius, that
may result from the effect of density waves being launched at
different Lindblad resonances (see eg. Goldreich & Tremaine
1979), we restarted this run after moving the super-Earth
to a new radius r = 2.5. It was eventually found to migrate
inwards but, when compared to the original rate, at a very
small and decreasing rate. This indicates the possibility of
either reaching ∼ r = 1.6 or stalling at some point further
out. At r ∼ 2, where inward migration starts to slow, the
surface density profile is affected by tidal effects due to the
giant planet and thus the migration of the super-Earth can
be affected. Finally we ran the same simulation but with the
giant planet mass taken to be two Jupiter masses, starting
with the same initial surface density profile but with the ini-
tial radius for the super-Earth taken to be r = 2.5. In this
case the initial inward migration stalls at r ∼ 2 after which
the planet undergoes slow outward migration which could
be followed for ∼ 1000 planet orbits.
We comment that in all of these cases that have sus-
tained interrupted inward migration over long periods of
time, we do not see any removal of the effect on an estimated
libration time scale for the horseshoe region, as might be an-
ticipated if corotation torques operated in an otherwise qui-
escent coorbital region. In this context we remark that pa-
rameters are comparable to those adopted in Paardekooper
& Papaloizou (2008) (see eg. their Fig. 21) with the libration
period being 100−200 planet orbits, which is much less than
1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 0 5000 10000 15000 20000 25000 30000semi-major axistime 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 0 10000 20000 30000 40000 50000semi-major axistime 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 0 2000 4000 6000 8000 10000 12000semi-major axistime12
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 10. The azimuthally averaged surface density profiles for an initial surface density distribution corresponding to
the standard disc (left) and for an initial surface density distribution corresponding to the more extensive disc (right). The
snapshots are taken after 3100 (black), 6200 (light grey) and 10000 (dark grey) the time units.
Figure 11. Orbital evolution for simulations starting with the more extensive surface density profile given by equation (5)
is illustrated. The left panel shows the evolution of the semi-major axis of a super-Earth in a presence of a gas giant with
one Jupiter-mass (black curve) and two Jupiter masses (grey curve). The initial value is 1.65 in each case. The right panel
shows the evolution of the semi-major axis of a planet of 1 Earth mass (light grey curve), 5.5 Earth masses (black curve)
and 10 Earth masses (dark grey curve) in a disc with aspect ratio h = 0.05. The initial value is 1.9 in each case.
the duration of our simulations. A similar situation applies
to the inviscid local simulations we have carried out (see the
discussion at the end of section 5.2).
4.4 The effect of migration of the giant planet on
the migration of the super-Earth
To illustrate the effect on the migration of the super-Earth of
allowing the giant planet to migrate, we present the results
of a simulation starting with the extended surface density
profile, aspect ratio h=0.05 and ν = 2 × 10−6. The mass of
the super-Earth was 5.5 Earth masses. Thus this run has
the same parameters as for a corresponding one presented
in the upper panels of Fig. 12, except that the Jupiter mass
planet is allowed to migrate in response to the disc torques
acting on it. Results from this simulation are illustrated in
Fig. 13.
If effects that interrupt inward migration are associated
with the neighbourhood of the giant planet, the super-Earth
should naturally tend to move with it. In this case, although
the giant moves from r = 1 to r = 0.8, the super-Earth still
moves slowly outwards. However, the ratio of its orbital ra-
dius to that of the giant reaches ∼ 2 where calculations pre-
sented above indicate that tides still operate. Thus the above
simulation is approximately consistent with the super-Earth
maintaining the same relationship with the giant planet as
when it does not migrate.
5
LOCAL SHEARING BOX SIMULATIONS
OF A PLANET EMBEDDED IN A DISC
WITH INDEPENDENTLY EXCITED
DENSITY WAVES
The results obtained in Section 4 for the migration of a
super-Earth may be complicated by the presence of a deep
gap opened by the gas giant and the particular way that
forces density waves in the disc.
To establish the fact that the essential features are a
planet embedded in a disc hosting outward propagating den-
sity waves, we decided to investigate the phenomenon by
performing local simulations with a planet embedded in a
disc with density waves independently forced by an inde-
pendent mechanism that did not involve additional planets
producing a very deep gap. These simulations may also be
performed at much higher resolution of the coorbital region
than the global ones and could be used to show that the qual-
itative nature of process was robust to changes in parameters
such as gravitational softening density wave amplitude and
applied viscosity.
5.1 Local Model
We adopt a local Cartesian coordinate system (x, y, z) uni-
formly rotating with angular velocity Ωp k and with origin
located at a point where the Keplerian angular velocity is
Ωp as seen in the inertial frame. The local radial coordinate
is x and y is the orthogonal coordinate pointing in the di-
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0.0008 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5surface densityradial distance from the star 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5surface densityradial distance from the star 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2 2.05 2.1 2.15 0 5000 10000 15000 20000 25000 30000 35000semi-major axistime 1.78 1.8 1.82 1.84 1.86 1.88 1.9 1.92 1.94 1.96 1.98 2 0 10000 20000 30000 40000 50000semi-major axistimeA super-Earth migrating in a disc with density waves excited by a gas giant
13
Figure 12. Upper panels: The semi-major axis of the super-Earth as a function of time for ν = 2 × 10−6 (lower black curve) and
ν = 10−5(upper gray curve) is plotted in the left panel. The azimuthally averaged surface density profiles, with the planet position
indicated, for ν = 2 × 10−6 (upper black curve) and ν = 10−5(lower grey curve) at t = 12000 are plotted in the right panel. Lower
panels: The left panel shows the semi-major axis of the super-Earth as a function of time for the simulations with ν = 0. Results are
presented for a one Jupiter mass giant with a super-Earth starting at r = 1.65 (black), r = 2.5 in a continuation run (light grey) and
a two Jupiter mass giant with a super-Earth starting at r = 2.5 (dark grey). The right panel shows the azimuthally averaged surface
density as a function of radius with the final positions of the planets indicated for the three cases.
Figure 13. Left: The evolution of the semi-major axes of the super-Earth (upper curve) and the Jupiter mass planet (lower
curve) when the latter is allowed to migrate. Right: The azimuthally averaged surface density profiles of the disc after 400
(black), 6000 (light grey) and 12000 (dark grey) time units with the positions of the super-Earth indicated.
rection of the shear. The governing hydrodynamic equations
are the continuity equation
= −∇ · (Σu),
∂Σ
∂t
and the equation of motion
to the planet is given by
Φp = −
GM2
(cid:112)x2 + y2 + 2
(cid:112)d2
GM2
max + 2
2
2
(6)
Φp = −
+ Φext, x2 + y2 < d2
max, and
+ Φext, x2 + y2 > d2
max.
(8)
∂u
∂t
+ u · ∇u + 2Ωp k ∧ u = 3Ω2
pxi − 1
Σ
∇P − ∇Φ + fν ,
(7)
where i is the unit vector in the x direction. The gravita-
tional potential, Φ, for these simulations, is taken to be due
to the single planet of mass M2, assumed to be fixed at the
origin of the rotating coordinate system, and the action of a
specified external forcing potential Φext. The potential due
The simulations we undertake are two dimensional with
no vertical dependence, and the gravitational potential is
chosen to be consistent with this. We remark that the indi-
rect term and self-gravity are neglected in this model and as
before 2 is the gravitational softening length for the planet
potential. For the simulations reported in detail here we
adopted 2 = 0.3H and dmax = 3H. The simulation set up
is similar to that described in Papaloizou et al. (2004) with
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
1.6 1.65 1.7 1.75 1.8 1.85 1.9 0 2000 4000 6000 8000 10000 12000semi-major axistime 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5surface densityradial distance from the star 1.5 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 2.5 2.6 0 5000 10000 15000 20000 25000 30000 35000 40000semi-major axistime 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0.0008 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5surface densityradial distance from the star 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 0 2000 4000 6000 8000 10000 12000 14000semi-major axistime 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5surface densityradial distance from the star14
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 14. Results are shown for the case with q = 0.1, C0 = 0.2, and no applied viscosity, (ν = 0). The upper left panel shows
surface density contours 48 orbits after initiation. The upper right panel shows the mean surface density profile averaged over the
y direction 56 orbits after initiation. The units are such that the uniform initial value for all local simulations was 0.001. The
central left panel shows force components in the y direction acting on the planet evaluated as a running time average. The units are
arbitrary but the same for all local simulations illustrated. The uppermost plot gives the contribution from x < 0. The lowermost
plot gives the contribution from x > 0. The central plot shows the total force component in the y direction acting on the planet
due to material in the computational domain. The central right hand panel indicates streamlines superposed on surface density
contours for the domain −3 < y/H < 3 and −0.6 < x/H < 0.6 48 orbits after initiation. The lower left panel indicates streamlines
superposed on surface density contours for the domain −6 < y/H < −3 and −0.6 < x/H < 0.6 56 orbits after initiation. The right
hand panel gives the corresponding plot for −3 < y/H < 3 and −0.6 < x/H < 0.6.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
15
Figure 15. Results are shown for the case with q = 0.1, C0 = 0.1, and ν/(H 2Ωp) = 0.02. The upper left panel shows surface
density contours 48 orbits after initiation. The upper right panel shows the mean surface density profile averaged over the y direction
48 orbits after initiation. The lower left panel shows force components in the y direction acting on the planet evaluated as a running
time average. The uppermost plot gives the contribution from x < 0. The lowermost plot gives the contribution from x > 0.
The central plot shows the total force component in the y direction acting on the planet due to material in the computational
domain. The lower right hand panel indicates streamlines superposed on surface density contours for the domain −3 < y/H < 3
and −0.6 < x/H < 0.6 48 orbits after initiation.
the adoption of the boundary condition that the solution be
periodic in shearing coordinates. The distance scale, dmax,
is used to flatten the planet potential for x2 + y2 > d2
max in
order to make it effectively periodic in shearing coordinates
when applied to the computational domain. The length scale
is H, which as for the global models, can be regarded as the
putative local disc semi-thickness. The equation of state is
thus strictly isothermal with P = ΣH 2Ω2
p. The viscous force
per unit mass, fν , is taken to be the standard Navier Stokes
form with kinematic viscosity ν as in the global simulations.
The planet mass appears only through specification of
the dimensionless ratio q ≡ (M2/M∗)/h3 ( Papaloizou et al.
2004) which has to be specified for each simulation. Because
all lengths are expressed in terms of H and times in units
of Ω−1
p , no other parameters need to be specified. However,
when M∗ is specified, the aspect ratio h ≡ H/R can be found
once, R, the orbital radius of the origin of the coordinate
system is in turn found from Kepler's law in the form R3 =
GM∗/Ω2
p.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
The computational domain was a rectangle of side 8H
in the x direction and 4πH in the y direction. We adopt
the convention that the positive x direction corresponds to
the outward or outer direction and the opposite direction
to the inward or inner direction. We have performed simu-
lations with the equally spaced grid resolutions (Nx, Ny) =
(261, 300) which we refer to as the standard resolution and
also (Nx, Ny) = (522, 600), which we refer to as higher res-
olution, in order to confirm convergence of the results. Cal-
culations were initiated with a uniform surface density and
velocity u = (0,−3Ωpx/2, 0) corresponding to local Kep-
lerian shear. The external forcing potential was applied in
order to excite density waves and so for these simulations
plays the role of the giant planet. For 0 < x − xmin < 2H,
we adopted the form
Φext =C0Ω2
pH 2 sin(π(x − xmin)/(2H))×
cos((y − ymin − vwavet)/(2H)),
(9)
16
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 16. Results are shown for the case with q = 0.05, C0 = 0.1, and ν/(H 2Ωp) = 0.02. The upper left panel shows force
components in the y direction acting on the planet evaluated as a running time average. The uppermost plot gives the contribution
from x < 0. The lowermost plot gives the contribution from x > 0. The central plot shows the total force component in the y
direction acting on the planet due to material in the computational domain. The upper right panel shows the mean surface density
profile averaged over the y direction 56 orbits after initiation. The lower left panel indicates streamlines superposed on surface
density contours for the domain −6 < y/H < −3 and −0.6 < x/H < 0.6 56 orbits after initiation. The right hand panel gives the
corresponding plot for −3 < y/H < 3 and −0.6 < x/H < 0.6.
where xmin and ymin are respectively the minimum values
of x and y in the computational domain, the propagation
speed vwave = (3π − 3/2)ΩpH and C0 is a scaling constant
specified for each simulation. For x − xmin > 2H no forcing
was applied. This is of the form of a harmonic forcing with
a single wavelength in the y direction while being localized
in the x direction. It is associated with a propagation speed
vwave in the y direction so that there is a pattern speed as
seen by the super-Earth which is at rest in the centre of
the computational domain. This plays the role of the cor-
responding pattern speed of the giant planet in the global
simulations. For the case we have adopted, the putative ma-
terial that would comove with the potential perturbation
is located at x = −(2π − 1)H, which is a distance 1.28H
interior to the inner boundary. Accordingly this mimics a
perturbation from an orbiting inner planet. The outer Lind-
blad resonance is located in the computational domain at a
distance 0.21H from the inner boundary. For the particular
choice of parameters we have made, density waves launched
from the outer Lindblad resonance propagate outwards. As
the inner Lindblad resonance is not present effects due to
inward propagating waves are minimized as, would be ex-
pected if the excitation was indeed due to an inner planet.
As the waves propagate across the whole domain without
further excitation for x − xmin > 2H, their dissipation is
optimized. This feature also limits the influence of distur-
bances from the periodic system of boxes that arise through
application of the shear periodic boundary conditions.
5.2 Local simulation results
We begin by considering a case with q = 0.1, C0 = 0.2,
and zero applied Navier-Stokes viscosity. The density pro-
file is the most strongly perturbed in this simulation which
was run for 400Ω−1
time units. Some results are shown in
Fig. 14. Surface density contours are shown 48 orbits after
initiation. These typically dominated by the trailing waves
excited by the potential Φext. The wake associated with the
p
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
17
embedded planet can scarcely be seen. The surface density
profile averaged over the y direction shows that the excited
waves causes momentum and mass transport that results in
material moving from inside the planet to outside the planet.
Tests have shown that this effect weakens if ν is increased
and/or the forcing amplitude C0 is decreased. We also plot
the running time averages of the force components acting
on the planet in the y direction. As expected, these show
that the contribution from exterior/interior to the planet is
decrease/increase the y momentum respectively. For these
runs the exterior contribution is always smaller resulting in
a net outward force on the planet. In the case of the simula-
tion described above, the net force is 60% of the one sided
interior force contribution. Note that these forces convert to
torques when scaled by the radius of the centre of the box.
Tests show the net force, though always outwards, decreases
in magnitude with wave forcing amplitude, planet mass and
gravitational softening parameter. It is also important to
remark that although the density response is dominated by
the forced waves the response of the planet is essential in
order to produce a net mean force. The specific mean force
is found to tend to zero with the planet mass. We also show
the streamlines superposed on surface density contours for
the domain −3 < y/H < 3 and −0.6 < x/H < 0.6 48 or-
bits after initiation in Fig. 14. In this case they indicate a
structure like that associated with normal horseshoes but
the planet is not precisely centered on a separatrix as would
be the case if pressure effects were negligible. This type of
structure was also seen in global simulations e.g. in the up-
permost right panel of Fig. 5 We remark that as for the
global simulations the streamline configuration varies with
time and so streamlines do not precisely indicate particle
trajectories. However, by following fluid elements we have
found that material initially with x < 0.25H remains close
to the planet such that x <∼ 0.5H in all simulations, which
indicates that the fluid particles do indeed turn. Another
streamline configuration we find is illustrated in the domain
−6 < y/H < 3 and −0.6 < x/H < 0.6 56 orbits after ini-
tiation in Fig. 14. In this case the flow is strongly affected
by the density waves, being deflected to move parallel to
shocks (see lower left panel). This has the consequence that
the flow through x = 0 is in many places reversed compared
to that expected for the usual horseshoe configuration, with
separatrix located far from the planet. This type of configu-
ration was also seen in the global simulations as in e.g. the
middle right panel of Fig. 4
In order to further illustrate the forces on an embed-
ded planet induced by independently excited density waves
we show results from two additional simulations. The first
was for the same parameters as above but with an applied
Navier-Stokes viscosity given by ν/(H 2Ωp) = 0.02. Results
for the simulation are presented in Fig. 15. In this case the
presence of the wakes due to the embedded planet are more
apparent in the surface density contours than the previous
case. The mean surface density profile averaged over the y
direction 48 orbits after initiation, though showing some os-
cillatory structure, is generally much flatter. The increased
value of ν has resulted in lower amplitude waves with less
associated angular momentum transport. The total mean
force component in the y direction acting on the planet
due to material in the computational domain is positive
but about seven times smaller than the previous simula-
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
tion again because of the smoothing action of viscosity. As
for the previous simulation, the streamlines in the domain
−3 < y/H < 3 and −0.6 < x/H < 0.6 48 orbits after ini-
tiation show indications of normal horseshoe turns but the
streamlines are distorted by the presence of the waves which
cause the planet to be no longer centered on a separatrix and
the flow to be parallel to shock fronts.
The second case we consider is the previous one but
now we reduce the planet mass to q = 0.05. The results are
shown in Fig. 16. In this case the running time average of
the total force component in the y direction acting on the
planet is positive but smaller than for the other simulations
we have discussed. The mean surface density profile aver-
aged over the y direction 56 orbits after initiation is flatter.
Interestingly the streamlines for the domain −6 < y/H < 3
and −0.6 < x/H < 0.6 56 at this time are qualitatively
similar to those for the case ν = 0 at the same time.
We remark that we have found, by following an ap-
propriate set of fluid elements, that there is no mean mass
flow through the coorbital region of the planet. Although
the planet is located on a rising surface density profile in
the case with ν = 0, the surface density structure is main-
tained by density waves rather than by either a mass flow,
and/or applied viscous stresses, as needed to maintain the
surface density structure in the planet trap of Masset et
al. (2006). In addition we note that, although coorbital ef-
fects are associated with the forces/torques acting on the
planet, it seems that the concept of saturation of corotation
resonances significantly affecting the positive force in the y
direction, plays no role. For the simulations presented here,
such effects should be manifest over a libration or return pe-
riod which corresponds to a time of 64 units for a distance
0.25H from the planet. The mean force behaviour is estab-
lished more quickly that this and remains for much longer
periods of time. The increase of the total force magnitude
that occurs on a long time scale when ν = 0 is due to the
secular effects associated with angular momentum transport
due to the waves.
6 THE MECHANISM LEADING TO
OUTWARD MIGRATION OF THE
SUPER-EARTH
In the previous sections we have described the results of nu-
merical experiments which indicate that outward migration
of a low mass planet may occur in a disc in the presence of
density waves excited by a forcing potential, which in the
global case was that due to a giant planet. We now consider
this process further with the aim of investigating its nature.
6.1 Migration of the super-Earth without the
source of density waves
The presence of a gas giant in the close internal orbit inte-
rior to that of the low mass planet affects the migration of
the low-mass planet halting or reversing its migration. This
occurs independently of the inclusion of the direct gravita-
tional interaction between the giant and the super-Earth.
Accordingly it must occur through modification of the sur-
face density profile through the action of density waves. If
the source of density wave excitation is removed, the low
18
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
mass planet should undergo type I migration which in our
case is directed inwards. In order to demonstrate this, be-
yond some point of some of our simulations we set the giant
planet mass to zero. The above expectations regarding mi-
gration are confirmed. We consider a simulation where the
non migrating giant planet mass was one Jupiter mass, the
disc aspect ratio was h = 0.05, the initial orbital radius of
the super-Earth was 1.9 and the initial surface density pro-
file was given by Eq.(5). The giant planet was removed at
time t = 750.
The form of the migration, with and without the re-
moval of the giant planet, is shown in Fig. 17. This clearly
indicates that, when the giant planet is removed, the super-
Earth starts to undergo inward migration at a rate similar
to that it had initially, whereas when it is retained outward
migration is ultimately attained. The right panel shows the
azimuthally averaged surface density profiles at the end of
both simulations.
The disc surface density contours in both these cases are
illustrated in Fig. 18 at time t = 795. These are very smooth
for the case where the giant planet was removed because
of the absence of density waves. Note too that in spite of
this, the inner cavity with a region of increasing azimuthally
averaged surface density, originally produced by the giant
remains and the presence this surface density structure does
not affect the inward migration of the low mass planet. This
is in contrast to the case when the giant is allowed to migrate
inwards, described in section 4.4 The edge profile in that
case adjusts in a similar way while outward migration is
maintained.
6.2 Angular momentum transport and dissipation
To explore the nature of the positive torques acting on the
super-Earth, we consider the relationship between angular
momentum flux and energy dissipation due to shocks aris-
ing from the density waves within a coorbital annulus. The
conservation of angular momentum integrated over azimuth
may be written in the form
ΣrT dϕ
(cid:90) 2π
(10)
+
=
∂ρj
∂t
∂Fj
∂r
0
Here the rate of angular momentum flow through a circle of
radius, r, is
Fj =
Σr2vrvϕdϕ,
(11)
(cid:90) 2π
(cid:90) 2π
0
the angular momentum per unit length is
Σr2vϕdϕ
0
ρj =
(12)
and T = Tgi + Tse is the torque per unit mass acting on the
disc which has contributions Tgi and Tse due to the giant and
super-Earth respectively. In the above, and also below, for
simplicity we neglect the viscous contribution to the fluxes
although including this would not affect our later discussion.
Assuming that the forcing due to the giant and super-
Earth has pattern speeds Ωi,0 and Ωp respectively, We may
also derive a conservation law in the form
∂ρJac
∂t
+
∂FJac
∂r
= −w + (Ωp − Ωi,0)
ΣrTsedϕ.
(13)
(cid:90) 2π
0
Here
ρJac = ρE − Ωi,0ρj,
and
FJac = FE − Ωi,0Fj,
with the energy flux given by
(cid:90) 2π
(cid:90) 2π
0
Σrvr
(cid:0)v2/2 + c2
Σr(cid:0)v2/2 + c2
s ln Σ + Φ(cid:1) dϕ,
s ln Σ + Φ(cid:1) dϕ.
and the energy per unit length given by
FE =
ρE =
(14)
(15)
(16)
(17)
0
Here, for simplicity we have adopted a strictly isothermal
equation of state as adopted in the local simulations. As the
coorbital region is of small radial extent this should be a
reasonable approximation for the global simulations also.
The rate of energy dissipated per unit length in the ra-
dial direction by the density waves is w. When w and Tse
are zero, on integration over the radial domain with van-
ishing boundary fluxes, Eq.(13) yields the Jacobi invariant
for the flow. Noting that ρJac is negative, we see that the
energy dissipation rate can be balanced either against a neg-
ative torque produced on the disc by the super-Earth, or by
a rate of increase in the magnitude of ρJac locally. The lat-
ter corresponds to torquing up the disc matter. Thus shock
dissipation can in part be associated with balancing of neg-
ative torque on the disc produced by the super-Earth which
implies a positive torque acting on and outward migration of
the super-Earth. Equation (10) also shows that in an approx-
imate steady state a decreasing outward angular momentum
flux can be balanced by a negative torque acting on the disc
due to the super-Earth.
We investigate the relationship between Fj and the
torque acting on the super-Earth during a simulation per-
formed for the disc with initial surface density profile given
by Eq.(5) and h = 0.05. The giant planet is one Jupiter mass
and the super-Earth is 5.5M⊕. The giant planet is not al-
lowed to migrate while the super-Earth is found to migrate
slowly outwards. In Fig. 19 we show the comparison of −Fj
with the torque acting on the planet over a time span just
exceeding one orbital period of the super-Earth after about
t = 11800. It is seen that these behave similarly.
We have calculated the torque acting on the planet due
to the disc matter located in the vicinity of the planet be-
tween r = 1.6 and r = 2.3, the planet being located at
around r = 1.89, at the moments of the time indicated by
the vertical lines in Fig. 19. In Fig. 20 we present the con-
tours of the surface density and in Fig. 21 we show the az-
imuthally averaged angular momentum flow rate through a
circle of radius r as a function of radius calculated according
to Eq.(11) at the same times. Note that, although the aver-
age value is positive, the torque values are small on account
of cancellation effects. The largest magnitude negative value
occurs at the first time on account of a large surface density
excess to the left of the planet. The largest positive value oc-
curs at the third time on account of a positive surface density
excess close to the planet. The azimuthally averaged angu-
lar momentum flow rate through a circle of radius r shows a
complicated temporal and spatial dependence on account of
the dynamic situation and the interaction of waves excited
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
A super-Earth migrating in a disc with density waves excited by a gas giant
19
Figure 17. The left panel shows the evolution of the semi-major axis of a low mass planet in the presence of a gas
giant (grey curve), and after the gas giant removal from the system (black curve). The right panel shows the azimuthally
averaged surface density profiles of the disc for the cases with (grey curve) and without (black curve) a gas giant at the end
of simulations (at time t = 8000).
Figure 18. The surface density contours in the disc with waves excited by a gas giant and a super-Earth migrating through
them (left) and the surface density distribution in the disc after a gas giant has been removed with a super-Earth migrating
in the resulting smooth surface density profile (right).
by the giant and the super-Earth. Nonetheless, we note that
Eq.(10) implies that provided Fj vanishes in the gap region,
−Fj(r = 2.3) should equal the sum of the torques acting on
both planets together with the rate of increase of angular
momentum of the disc material (note that this can actually
be negative in some places as material does seep across the
gap). ¿From Fig. 19 we see that this quantity is of a similar
form as the torque on the super-Earth but about three times
larger, consistent with the view that some fraction of the an-
gular momentum flow is transmitted to the super-Earth.
7 CONSEQUENCES OF THE WAVE-PLANET
INTERACTION FOR THE RESONANCE
CAPTURE OF A SUPER-EARTH BY A GAS
GIANT
A consequence of the results presented above is that outward
migration induced by density waves can prevent an initially
inwardly migrating super-Earth from reaching a 2:1 com-
mensurability with the giant. The angular momentum ex-
change during the super-Earth passage through the density
wave field induces its outward migration. It is thus prevented
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
from approaching close enough to the gas giant's orbit for
the commensurability to be reached. From this we expect
that in a system with a gas giant we could expect a low
mass planet to be in an exterior orbit, possibly close to, but
not exactly in 2:1 resonance. There are two observations
that relate to our studies. The first is the inference from
transit timing variations of a low mass companion of 15M⊕
that is close to an exterior 2:1 commensurability with the
giant planet (≈ 2MJ ) in the Wasp-3 system (Maciejewski et
al. 2010). However, further observations are needed to con-
firm this configuration. The other example is found in the
GJ876 system where the two outermost planets are close to
2:1 commensurability. The planets have masses 2.27MJ and
14.6M⊕ so they are within the regime where the mechanism
described in this paper can operate. The situation in GJ876
is complex because apart from the configuration of the outer
two planets, there is another inner giant in the system with
all three planets being in a Laplace resonance. Here we will
concentrate on the outer two planets: c and e in the GJ876
system. In order to perform simulations, we scale up the
masses of the outer two planets to be 7 Jupiter masses and
43 Earth masses so that with a central star of one solar
1.8 1.82 1.84 1.86 1.88 1.9 1.92 0 1000 2000 3000 4000 5000 6000 7000 8000semi-major axistime 0 0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5surface densityradial distance from the star20
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 19. The torque acting on the super-Earth (solid black line),
the angular momentum flow rate (−Fj ) at the position of the super-
Earth, across its orbital radius, averaged over azimuth (grey) and
−Fj (r = 2.3) averaged over azimuth (dotted). The vertical lines indi-
cate moments of time at which we show the surface density and the
azimuthally averaged angular momentum flux as a function of radius
below.
Figure 20. The surface density contours for the first six moments of time progressing from left to right ,upper then lower,
marked by vertical lines in Fig. 19.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
-0.0001-5e-05 0 5e-05 0.0001 0.00015 0.0002 11808 11810 11812 11814 11816 11818 11820torquetime1234567A super-Earth migrating in a disc with density waves excited by a gas giant
21
Figure 21. The azimuthally averaged angular momentum flow rate across a circle of radius, r, against radius at the first
six moments of time progressing left to right, upper then lower, marked by vertical lines in Fig. 19.
Figure 22. The left panel shows the evolution of the semi-major axis of a 43M⊕ planet orbiting exterior to a 7 Jupiter
mass gas giant. The initial orbital radius of the super-Earth is r = 1.9 (grey curve) and r = 1.65 (black curve), respectively.
The right panel shows the evolution of the semi-major axis of a 43M⊕ (grey curve) and a 20M⊕ planet when the giant
planet mass was reduced to one Jupiter mass (black curve).
mass, as adopted here, the mass ratios are the same as in
the GJ876 system.
In Fig. 22 (left panel) we show the evolution of the semi-
major axis of the outermost planet in a disc with aspect
ratio h = 0.05. In this particular case, the gas giant was
not allowed to migrate and the planets do not interact with
each other gravitationally. The initial surface density profile
was given by Eq.(5). The outermost planet was started in
an exterior orbit outside the 2:1 resonance. The evolution
proceeds differently for different initial locations in the disc,
but the final outcome of the evolution is similar. The smaller
planet always ends up at the same radial location, exterior
to the 2:1 commensurability with semi-major axis exceeding
twice that of the giant. This is illustrated in Fig. 22 for initial
locations of the planet of r = 1.65 and r = 1.9.
In order to explore this issue further, in Fig. 22 (right
panel) we compare the behaviour of 43M⊕ and 20M⊕ plan-
ets in the same initial disc but in the presence of a lower mass
gas giant with a mass of 1 Jupiter mass. The Jupiter mass
planet launches weaker density waves than the 7 Jupiter
mass planet. This enabled the 43M⊕ planet to pass through
the 2:1 resonance. Thus the density waves are too weak to
prevent the attainment of the commensurability in this case.
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
However the 20M⊕ planet stopped its migration just exte-
rior to resonance which indicates that smaller mass planets
do not reach the commensurability.
Finally, the above simulation for the 43M⊕ planet was
repeated with the gravitational
interaction between the
planets included so that capture into the 2:1 resonance takes
place. This is illustrated in Fig. 23 where we show the evo-
lution of the semi-major axis of the low mass planet, both
resonant angles (see e.g. Papaloizou & Szuszkiewicz 2005,
for definitions) and the difference in the mean longitudes.
Thus resonance capture is only possible for sufficiently
low mass giants that are not efficient enough as a source
of density waves to halt the inward migration. In order to
obtain a resonant configuration for a more massive giant,
we slowly increased its mass. However, this had to be done
after removal of significant amounts of disc material that
could act as a carrier of density waves. For example after
about 2300 time units we started to increase the mass of
the gas giant as a linear function of time until it reached the
mass of 7 MJ at a time t = 3341. At the beginning of this
procedure we instantly modified the surface density profile
so as to increase the size of the gap around the giant by
-0.0001-8e-05-6e-05-4e-05-2e-05 0 2e-05 4e-05 6e-05 8e-05 0.0001 0.00012 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star-0.00015-0.0001-5e-05 0 5e-05 0.0001 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star-0.0002-0.00015-0.0001-5e-05 0 5e-05 0.0001 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star-0.0001-8e-05-6e-05-4e-05-2e-05 0 2e-05 4e-05 6e-05 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star-0.00015-0.0001-5e-05 0 5e-05 0.0001 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star-0.0002-0.00015-0.0001-5e-05 0 5e-05 0.0001 1.6 1.7 1.8 1.9 2 2.1 2.2 2.3angular momentum fluxradial distance from the star 1.5 2 2.5 3 3.5 4 4.5 0 1000 2000 3000 4000 5000 6000semi-major axistime 1.4 1.5 1.6 1.7 1.8 1.9 2 0 1000 2000 3000 4000 5000 6000 7000 8000semi-major axistime22
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Figure 23. The evolution of the semi-major axis of a 43M⊕ planet (upper left panel), both resonant angles (lower panels),
and the difference in the mean longitudes (upper right panel) in the presence of the gas giant. After 2300 time units the
mass of the giant planet was increased from one to 7 Jupiter masses. The oscillations of the semi-major axis and one of the
resonant angles was found to increase in amplitude, and the other angle switches from libration to circulation.
adopting the following modification:
Σ → Σ × (r/2.5)90
Σ → 0.01Σ for
for
Σ → Σ for
r < 2.4
2.4 (cid:54) r < 2.5
r (cid:62) 2.5
(18)
After this has been carried out, the lower mass planet be-
comes located within an extensive gap, where there is far
less material, so that density waves become ineffective at
causing it to migrate outwards. Once we started to increase
the mass of the gas giant the oscillations of its semi-major
axis increased in amplitude and one resonant angle moved
from libration to circulation, as can be seen in Fig. 23. How-
ever the other resonant continues to librate, but with larger
amplitude. These details may depend on the detailed pro-
cedure we followed. Nonetheless, the planet remained close
to the 2:1 resonance as is observed in the Gliese 876 system
(Rivera et al. 2010).
However, the above scenario requires the giant to ac-
crete from a reservoir of material, possibly an interior disc,
after having removed most of the nearby outer disc. This
would appear unlikely. It may be that the lower mass planet
formed after the gas disc had dispersed and/or underwent
planetesimal migration (e.g. Raymond et al. 2009) to bring
it into a 2:1 commensurability from a larger radius.
8 CONCLUSIONS
In this paper we have studied a new mechanism that can re-
verse the type I migration of a low mass planet that would
occur in an unperturbed locally isothermal gaseous disc.
This mechanism operates when there is a source of trailing
density waves that propagate through the disc. The most
natural example of such source arises from the gravitational
perturbation by another planet and for global simulations
we have focused on the case when a gas giant is present.
We have also carried out local simulations in a shearing box
where the forcing was due to an imposed harmonically vary-
ing potential with varying amplitude. In all cases the den-
sity waves produce shocks which are associated with angular
momentum transfer to the disc material. Coorbital disc ma-
terial then transfers angular momentum to the low mass
planet in the same direction, when it is scattered by it. This
is analogous to what happens with coorbital material un-
dergoing a slow drag in horseshoe orbits in the case of zero
pressure particle dynamics. However, in the cases considered
here the dynamics of the coorbital material is not station-
ary in an appropriately rotating frame and generally more
complex. In particular, we stress that although coorbital dy-
namics plays a role, the situation is unlike that pertaining to
standard corotation torques in initially unperturbed discs.
For example, because the dissipation and effective drag are
produced by shocks, which are viscosity independent and
the propagating density waves provide an external source of
angular momentum for the coorbital zone, issues of torque
saturation become irrelevant. Furthermore although there is
an interior gap, the mechanism described here depends on
the same source of angular momentum, rather than applied
viscous stresses or an accelerating inward accretion flow, to
maintain the surface density distribution and in this way it
differs from the trap mechanism of Masset et al. (2006).
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
1.54 1.56 1.58 1.6 1.62 1.64 1.66 0 2000 4000 6000 8000 10000 12000semi-major axistime 0 1 2 3 4 5 6 0 2000 4000 6000 8000 10000 12000λ1-2λ2time 0 1 2 3 4 5 6 0 2000 4000 6000 8000 10000 120002λ2-λ1-ω2time 0 1 2 3 4 5 6 0 2000 4000 6000 8000 10000 120002λ2-λ1-ω1timeA super-Earth migrating in a disc with density waves excited by a gas giant
23
The effect of the interaction with the density waves is
that the migration of a low-mass planet located in an ex-
terior orbit relative to the gas giant can be slowed down
and finally reversed. Thus we found that a planet with mass
in the super-Earth range cannot approach a Jupiter mass
planet close enough in order to form first order mean-motion
resonances with it. The migration was found to halt with
semi-major axis ranging between 1.6 and 2 times that of the
giant. Only when the low mass planet exceeded ∼ 40M⊕
was it able to attain a 2:1 commensurability. As the giant
planet mass is increased, even larger low mass planets would
be required.
Our results indicate that migrating the outermost
planet in the GJ876 system to its observed 2:1 commen-
surability through planet interaction with the gaseous disc
alone would be problematic for the reasons outlined above.
This may indicate that migration induced by planetesimals
after the clearance of the gas disc may have been significant
in that case.
ACKNOWLEDGMENTS
This work has been partially supported by NSC Grant
No. N N203 583740 (2011-2012) and MNiSW PMN grant
- ASTROSIM-PL "Computational Astrophysics. The for-
mation and evolution of structures in the universe: from
planets to galaxies" (2008-2011). We acknowledge support
from the Isaac Newton Institute programme "Dynamics of
Discs and Planets". Part of this research was performed dur-
ing a stay at the Kavli Institute for Theoretical Physics,
and was supported in part by the National Science Foun-
dation under Grant No. PHY05-51164. The simulations re-
ported here were performed using the HAL9000 cluster of
the Faculty of Mathematics and Physics of the University of
Szczecin. JCBP acknowledges support through STFC grant
ST/G002584/1. We wish also to thank Adam (cid:32)Lacny for his
helpful comments. Finally, we are indebted to Franco Fer-
rari for his continuous support in the development of our
computational techniques and computer facilities.
REFERENCES
Adams, F. C., Bloch, A. M., 2009, ApJ, 701, 1381
Artymowicz, P., Clarke, C. J., Lubow, S. H., & Pringle, J.
E. 1991, ApJ, 370, L35
Balbus, S. A., Papaloizou, J. C. B., 1999, ApJ, 521, 650
Baruteau, C., Masset, F., 2008, ApJ, 672, 1054
D'Angelo. G., Lubow, S. H., Bate, M. R., 2006, ApJ, 652,
1698
De Val-Borro, M., Artymowicz, P., D'Angelo, G., Peplinski,
A., 2007, A&A, 471, 1043
Dermott, S.F., Murray, C.D., Sinclair, A.T., 1980, Nature,
284, 309
Goldreich, P., & Tremaine, S. , 1979, ApJ, 233, 857
Hasegawa, Y., Pudritz, R. E., 2010, ApJ, 710, L167
Johnson, E. T., Goodman, J. & Menou, K. 2006, ApJ, 647,
1413
Kley, W., Bitsch, B., & Klahr, H. 2009, A&A, 506, 971
Kley, W., Crida, A., 2008, A& A, 487, L9
Kley, W., Dirksen, G., 2006, A&A, 447, 369
Laughlin, G., Chambers, J., Fischer, D., 2002, ApJ, 579,
455
Laughlin, G., Steinacker, A., Adams, F.C., 2004, Astro-
phys. J. 608, 489
Li, H., Lubow, S. H., Li, S., Lin, D. N. C., 2009, ApJL,
690, L52
Lin, D.N.C., Bodenheimer, P., Richardson, D.C., 1996, Na-
ture, 380, 606
Lynden-Bell, D., Kalnajs, A.J., 1972, MNRAS, 157, 1
Maciejewski, G., et al. 2010, MNRAS, 407, 2625
Masset, F. S., 2000, A&AS, 141, 165
Masset, F. S., Morbidelli, A., Crida, A., Ferreira, J., 2006,
ApJ, 642, 478
Masset, F. S., Papaloizou, J. C. B., 2003, ApJ, 588, 494
Matsumura, S., Pudritz, R. E., Thommes, E. W., 2005,
in Protostars and Planets V, Proceedings of the Confer-
ence held October 24-28, 2005, in Hilton Waikoloa Village,
Hawai'i. LPI Contribution No. 1286., p.8544
Menou, K., Goodman, J., 2004, ApJ, 606, 520
Nelson R. P., Papaloizou J.C. B., Masset F., Kley W., 2000,
MNRAS, 318, 18
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
24
E. Podlewska-Gaca, J.C.B Papaloizou and E. Szuszkiewicz
Nelson R. P., Papaloizou J.C. B., 2002, MNRAS, 333, L26
Nelson R. P., Papaloizou J.C. B., 2004, MNRAS, 350, 849
Ogihara, M., Duncan, M. J., Ida, S., 2010, ApJ, 721, 1184
Ogilvie, G. I., & Lubow, S. H. 2003, ApJ, 587, 398
Paardekooper, S.-J.; Mellema, G., 2006, A& A, 459, L17
Paardekooper, S.-J.; Papaloizou, J. C. B., 2008, A&A, 485,
877
Paardekooper, S.-J.; Papaloizou, J. C. B., 2009, MNRAS,
394, 2283
Paardekooper, S.-J.; Papaloizou, J. C. B., 2009, MNRAS,
394, 2297
Paardekooper, S.-J., Baruteau, C., Crida, A., Kley, W.,
MNRAS, 401, 1950
Paardekooper, S.-J., Baruteau, C., Kley, W., MNRAS, 410,
293
Papaloizou, J. C. B., 2002, A&A, 388, 615
Papaloizou, J. C. B., Larwood, J, 2000, MNRAS,315, 823
Papaloizou, J. C. B., Nelson, R. P., & Masset, F. 2001,
A&A, 366, 263
Papaloizou, J.C.B., Nelson, R.P. & Snellgrove, M.D., 2004,
MNRAS, 350, 829
Papaloizou, J. C. B., Szuszkiewicz, E., 2005, MNRAS, 363,
153
Papaloizou, J. C. B., Szuszkiewicz, E., 2010, in Extrasolar
Planets in Multi-Body Systems: Theory and Observations,
K. Go´zdziewski, A. Niedzielski and J. Schneider (eds),
EAS Publication Series, 42, 333
Pierens, A., Nelson, R. P., 2008, A&A, 482, 333
Podlewska, E., Szuszkiewicz, E., 2008, MNRAS, 386,1347
Podlewska, E., Szuszkiewicz, E., 2009, MNRAS, 397,1995
Raymond, S.N., Armitage , P. J. & Gorelick, N., 2009,
ApJL, L88
Rivera, E., Laughlin, G., Butler, P., Vogt, S., Haghigh-
ipour, N. & Meschiari, S., 2010, ApJ, 719, 890
Tanaka, H., Takeuchi, T., Ward, W. R., 2002, ApJ, 565,
1257
Terquem, C., 2003, MNRAS, 341, 1157
Thommes, E. W., 2005, ApJ, 626, 1033
Ward, W. R., 1997, Icarus, 126, 261
Yamada, K., Inaba, S., 2011, MNRAS, 411, 184
Yu, C., Li, H., Li, S., Lubow, S. H., Lin, D. N. C., 2010,
ApJ, 712, 198
Ziegler U., 1998, Comp. Phys. Commun., 109, 111
c(cid:13) 2010 RAS, MNRAS 000, 1 -- 24
|
1506.01402 | 2 | 1506 | 2015-09-10T19:19:46 | Was Comet C/1945 X1 (du Toit) a Dwarf, SOHO-Like Kreutz Sungrazer? | [
"astro-ph.EP"
] | The goal of this investigation is to reinterpret and upgrade the astrometric and other data on comet C/1945 X1, the least prominent among the Kreutz system sungrazers discovered from the ground in the 20th century. The central issue is to appraise the pros and cons of a possibility that this object is --- despite its brightness reported at discovery --- a dwarf Kreutz sungrazer. We confirm Marsden's (1989) conclusion that C/1945 X1 has a common parent with C/1882 R1 and C/1965 S1, in line with the Sekanina-Chodas (2004) scenario of their origin in the framework of the Kreutz system's evolution. We integrate the orbit of C/1882 R1 back to the early 12th century and then forward to around 1945 to determine the nominal direction of the line of apsides and perform a Fourier analysis to get insight into effects of the indirect planetary perturbations. To better understand the nature of C/1945 X1, its orbital motion, fate, and role in the hierarchy of the Kreutz system as well as to attempt detecting the comet's possible terminal outburst shortly after perihelion and answer the question in the title of this investigation, we closely examined the relevant Boyden Observatory logbooks and identified both the photographs with the comet's known images and nearly 20 additional patrol plates, taken both before and after perihelion, on which the comet or traces of its debris will be searched for, once the process of their digitization, currently conducted as part of the Harvard College Observatory's DASCH Project, has been completed and the scanned copies made available to the scientific community. | astro-ph.EP | astro-ph |
Version July 10, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
WAS COMET C/1945 X1 (DU TOIT) A DWARF, SOHO-LIKE KREUTZ SUNGRAZER?
1Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, U.S.A.
2Ostlandring 53, D-25335 Elmshorn, Schleswig-Holstein, Germany
Zdenek Sekanina1 & Rainer Kracht2
Version July 10, 2018
ABSTRACT
The goal of this investigation is to reinterpret and upgrade the astrometric and other data on comet
C/1945 X1, the least prominent among the Kreutz system sungrazers discovered from the ground in
the 20th century. The central issue is to appraise the pros and cons of a possibility that this object is
-- despite its brightness reported at discovery -- a dwarf Kreutz sungrazer. We confirm Marsden's
(1989) conclusion that C/1945 X1 has a common parent with C/1882 R1 and C/1965 S1, in line with
the Sekanina-Chodas (2004) scenario of their origin in the framework of the Kreutz system's evolution.
We integrate the orbit of C/1882 R1 back to the early 12th century and then forward to around 1945
to determine the nominal direction of the line of apsides and perform a Fourier analysis to get insight
into effects of the indirect planetary perturbations. To better understand the nature of C/1945 X1, its
orbital motion, fate, and role in the hierarchy of the Kreutz system, as well as to attempt detecting the
comet's possible terminal outburst shortly after perihelion and answer the question in the title of this
investigation, we closely examined the relevant Boyden Observatory logbooks and identified both the
photographs with the comet's known images and nearly 20 additional patrol plates, taken both before
and after perihelion, on which the comet or traces of its debris will be searched for, once the process of
their digitization, currently conducted as part of the Harvard College Observatory's DASCH Project,
has been completed and the scanned copies made available to the scientific community.
Subject headings: comets: general -- methods: data analysis
1. INTRODUCTION
As the most extensive system of genetically related
comets in existence, the Kreutz sungrazers represent an
inexhaustible source of research opportunities. Kreutz's
(1888, 1891, 1901) celebrated orbital studies described
the motions of the early bright members, showing that
they moved about the Sun in similar, extraordinarily
elongated paths, with orbital periods of up to about 1000
yr, yet approaching the Sun's surface to within one solar
radius at perihelion. Kreutz's work was followed by many
more studies in the 20th century, with those by Marsden
(1967, 1989) standing out as the most important.
Because all Kreutz sungrazers are fragments of one pro-
genitor, their orbits' lines of apsides are nearly perfectly
aligned; the scatter is only a small fraction of 1◦, a prod-
uct of indirect perturbations by the planets, Jupiter in
particular (Marsden 1967), and of the process of cas-
cading fragmentation (Sekanina 2002). The research on
the Kreutz system has recently accelerated explosively
thanks to vast new evidence provided by imagers on
board the spacecraft dedicated to the exploration of the
Sun, especially the coronagraphs C2 and C3 of the Solar
and Heliospheric Observatory (SOHO ; see Brueckner et
al. 1995) and the coronagraphs COR2 and imagers HI1 of
the Solar Terrestrial Relations Observatory's two probes
(STEREO-A and B ; see Howard et al. 2008). Over the
past two decades, these instruments allowed detection, in
close proximity of the Sun, of thousands of minor Kreutz
system's members, referred to hereafter as the dwarf sun-
grazers, which keep streaming toward the Sun but always
disintegrate shortly before reaching perihelion.
Electronic address: [email protected]
[email protected]
The directions of the lines of apsides of 1600 dwarf
Kreutz sungrazers derived from their published gravi-
tational orbits were recently shown by us (Sekanina &
Kracht 2015; hereafter referred to as Paper 1) to be dis-
tributed along an arc of 25◦ (sic!) in the ecliptical lat-
itude, failing utterly to comply with the condition of
directional alignment. We determined that this major
effect was due to a neglected nongravitational accelera-
tion in the dwarf sungrazers' motions, which was orders
of magnitude greater than the nongravitational acceler-
ations in the motions of the cataloged comets in nearly-
parabolic orbits with perihelia a few tenths of AU from
the Sun or more, topping in exceptional cases the Sun's
gravitational acceleration.
In summary, the preperihe-
lion disintegration and a very high erosion-driven non-
gravitational acceleration of the orbital motion are two
fundamental attributes of the dwarf Kreutz sungrazers.
2. SUNGRAZING COMET C/1945 X1 (DU TOIT)
On 1945 December 11, a comet was discovered photo-
graphically by D. du Toit at the Harvard College Obser-
vatory's Boyden Station near Bloemfontein, South Africa
(Paraskevopoulos 1945); nowadays this comet is referred
to as C/1945 X1. The object moved rapidly toward the
Sun and its brightness at discovery was reported as mag-
nitude 7. Additional plates were taken at Boyden on the
following nights until December 15, but the five estimated
astrometric positions were not communicated until 1946
January 2, when Cunningham (1946a) used them to com-
pute three very preliminary parabolic orbits. They indi-
cated that the comet was apparently a Kreutz sungrazer
that had passed perihelion five days before the cable was
sent. Cunningham (1946b) pointed out that his search
ephemeris, based on one of the three orbits, was uncer-
2
Sekanina & Kracht
tain "by many degrees." Published accounts show that
after December 15 the comet was lost and never seen
again. Contrary to expectations, it did not become a
brilliant object near and/or after perihelion.
Its fail-
ure to develop a bright headless tail for a limited period
of time after perihelion -- contrary to such sungrazers
as C/1887 B1 (see the references in Sekanina 2002) and
C/2011 W3 (e.g., Sekanina & Chodas 2012) -- suggests
that C/1945 X1 may have disintegrated already before
perihelion, as do the dwarf Kreutz sungrazers. This pos-
sibility raises a question of whether or not this object
was the only dwarf Kreutz sungrazer discovered and re-
peatedly observed from the ground outside a total solar
eclipse.1 To address this issue in detail requires that
three critical points be answered:
(1) Can a very high nongravitational acceleration be
conclusively detected in the comet's orbital motion from
the five Boyden observations?
(2) Why was the comet so bright 17 days before perihe-
lion? Was it in outburst or was the reported brightness
grossly overestimated?
(3) Was the absence of the comet's relics after perihe-
lion conditioned on a compelling qualifier or constraint,
so that the comet's preperihelion disintegration could be
subject to doubt?
We present new evidence in the following sections that
allows us to comment on these points and to chart the
lines of attack in the near future.
3. THE BOYDEN PHOTOGRAPHIC OBSERVATIONS
A striking feature of the information on the comet's
Boyden observations is an extremely slow and protracted
progress in propagating their results, with an essentially
complete lack of details. This is most surprising, given
that C/1945 X1 was the first Kreutz sungrazer of the
20th century, after a pause of nearly 60 years.
An emphatic example of the slow progress is the pub-
lication of the comet's astrometry from the Boyden
plates. After a delayed message of the estimated posi-
tions (whose observation times were announced with a
precision to 1 hr!), there was no follow-up report and no
computation of an improved orbit until 22 years later,
when Marsden (1967), revealing that the plates were
measured and reduced by A. G. Mowbray in 1952, de-
rived several sets of parabolic elements. Strangely, these
astrometric data were reported by neither Mowbray him-
self nor L. E. Cunningham, for whom Mowbray was then
working (Hockey 2009). In fact, the positions were pub-
lished only 37 years after they were measured and re-
duced and 44 years after they were taken by the Boyden
observers (Marsden 1989)!
The circumstances of the comet's observations at Boy-
den have never been published. We were especially in-
terested in the telescopes or cameras employed to make
these photographic observations, in exposure times used,
and in the type of tracking (sidereal or on the comet) of
the post-discovery photographs. The only information
learnt from the literature was that in the 1940s Boyden
observers, such as M. J. Bester, discovered their comets
1 A dwarf sungrazer C/2008 O1 was detected during a total solar
eclipse on some exposures after a search based on data from SOHO
(Pasachoff et al. 2009). C/1882 K1 (Tewfik), discovered ∼3.5 hr
preperihelion at 5.6 R⊙ (Marsden 1967, 1989) during a total solar
eclipse on 1882 May 17, may have been a dwarf sungrazer as well.
with either the Metcalf 25-cm f/4.9 Triplet refractor or
the Bache 20-cm f/5.7 Doublet refractor (Cooper 2003,
2005) while examining plates for image quality.2
As integral part of the Harvard College Observatory's
Digital Access to a Sky Century@Harvard (DASCH) Proj-
ect, all Boyden plates are in the process of being dig-
itized, with the scans gradually made available to the
scientific community (Simcoe et al. 2006; Grindlay et al.
2012).3 The products of this effort will eventually play
a major role in our quest for information on the comet.
Of imminent interest to us are the observing logbooks
on the Harvard website; after some search, we have al-
ready been able to identify all five plates measured and
reduced by Mowbray. Three surprises surfaced: (i) the
comet was not discovered on a plate taken with the
Metcalf or Bache telescope; (ii) a never-reported post-
perihelion search was attempted with two instruments on
1946 January 8, 11 days after perihelion and the same
day that Cunningham's (1946b) ephemeris was issued;
and (iii) none of the follow-up observations was made by
the discoverer.
Table 1 presents the results of our search for the rel-
evant Boyden plates. For each, the individual columns
list: the plate number, the UT time of mid-exposure; the
approximate equatorial coordinates of the plate center
(converted to the equinox J2000); the local sidereal times
of the exposure's start and termination; the resulting ex-
posure time; the logbook reference; and the observer.
Three instruments were employed to observe, or search
for, the comet: the Metcalf Triplet (plates MF; a plate
scale of 167′′.3 mm−1) and two small cameras mounted
piggyback on the Metcalf, the Cooke 3.8-cm f/8.9 lens
(plates AM, including the discovery one; a plate scale
of 610′′.8 mm−1) and the Ross-Fecker 7.5-cm f/7 camera
(plates RB; a plate scale of 395′′.5 mm−1). The limit-
ing magnitudes of the three instruments are listed on
the Harvard website as, respectively, 17, 13 -- 14, and 15,
all much fainter than the comet's reported magnitude at
discovery.
All plates employed for C/1945 X1 had apparently a
blue sensitive emulsion (class L; see the DASCH website
footnote) and all three instruments used a plate size of
20.3 cm (in right ascension) by 25.4 cm (in declination),4
covering fields of 15◦.1 (Metcalf), 35◦.7 (Ross-Fecker), and
55◦.2 (Cooke) in angular extent along a diagonal.
The logbooks show that, of the five observations astro-
metrically measured and reduced by Mowbray and pub-
lished by Marsden (1989), the ones on December 11 -- 12
were made with the Cooke lens and only the ones on De-
cember 13 -- 15 with the Metcalf refractor. The logbooks
further show that the comet was ∼12◦ from the center
on the discovery plate, 2◦.9 on the December 12 Cooke
plate, always less than 1◦.2 on the Metcalf plates, and
just about 9◦.6 on the December 14 Cooke plate, which
apparently has never been measured. The question of
2 These plates were taken on behalf of H. Shapley, Director of the
Harvard College Observatory, for the purpose of studying southern
variable stars (Cooper 2003, 2005; van Heerden 2008).
3 The DASCH Project, at http://dasch.rc.fas.harvard.edu/,
involves digitization of more than 500 000 Harvard plates and is cur-
rently in progress; see also http://tdc-www.harvard.edu/plates/.
4 The website http://tdc-www.harvard.edu/plates/mf/ speci-
fies that Metcalf plates have generally sizes 8 in by 10 in, 14 in by
17 in, and, some early ones, 10 in by 12 in.
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
3
Boyden Plates with Reported Images or Possible Images of Comet C/1945 X1.
Table 1
Plate
numbera
UT time at
mid-exposure
R.A.
Dec.
start
stop
time
(min)
Logbook
reference
Observerc
Plate centerb
Sidereal time
Exposure
h
m
h
m
h
m
1945 Dec. 11.04709
12.04713
13.07210
14.03336
14.05898
15.06906
1946 Jan. 8.06517
8.07729
AM 25201d
AM 25206
MF 34987
AM 25208e
MF 34991
MF 34994
MF 35030
RB 14184
a AM plates taken with the Cooke lens; MF plates with the Metcalf telescope; and RB plate with the Ross-Fecker camera.
b Equinox J2000.
c The identity of the observer whose abbreviation in the logbooks was J could not be determined; see also Table 12.
d Discovery plate.
e This plate appears to have never been astrometrically measured and reduced.
am45b 0158
am45b 0158
mf25b 0086
am45b 0158
mf25b 0086
mf25b 0088
mf25b 0094
56
8
9 00
9 10
8 23
8 55
9 06
10 50
11 05
36
13
15 18.7
16 19.0
16 40.9
16 42.9
17 02.6
17 17.5
17 17.5
.7 −60◦31′
−65 22
−63 03
−70 12
−61 12
−58 39
−47 19
−47 19
26
7
7 30
8 40
7 43
8 25
8 51
10 05
10 25
90
90
30
40
30
15
45
40
rb10 148
du Toit
Bester
Britz
J(?)
J(?)
J(?)
Bester
Bester
whether or not the comet is located on the plates taken
on January 8 can only be answered after our orbital anal-
ysis (see Sections 7.1 and 8.2).
4. HISTORY OF THE COMET'S ORBIT DETERMINATION
The sets of orbital elements for comet C/1945 X1 com-
puted by Cunningham (1946a) were much too uncertain
to use as a basis of subsequent research. A more accu-
rate orbit was a gravitational parabolic approximation
derived by Marsden (1967), who employed the astrome-
try by Mowbray. Marsden concluded that the December
Bπ
+36◦.0
+35◦.5
+35◦.0
LINES OF APSIDES OF SOME
KREUTZ SUNGRAZERS
③
C/1945 X1
C/1843 D1
❩❩⑦
C/1963 R1
C/1882 R1
③
③
③
C/1880 C1
∗
PERT.
③
③C/1965 S1
③C/2011 W3
③
C/1970 K1
282◦.5
283◦.0
283◦.5
LONGITUDE OF PERIHELION, Lπ
Figure 1. Line of apsides for seven bright Kreutz sungrazers and
C/1945 X1. The plot of the perihelion longitude Lπ against per-
ihelion latitude Bπ shows that C/1945 X1, represented by Mars-
den's cataloged gravitational orbit, deviates significantly from the
cluster of the bright sungrazers, whose lines of apsides are closely
aligned. The asterisk marked Pert. is the apsidal-line position for
C/1945 X1 on the assumption that the comet is a fragment of a
common parent with C/1882 R1, the difference being due entirely
to the indirect planetary perturbations. For six comets the errors
of the coordinates Lπ, Bπ are smaller than the size of the symbols;
the estimated errors for C/1880 C1 and C/1945 X1 are depicted.
14 data point (which we found was measured on a MF
plate) was inconsistent with the other four positions and
he omitted it from what he described as the final set
of orbital elements obtained by least squares. The four
employed positions were fitted with a mean residual of
±2′′.1, the orbit appeared to be similar to those of C/1882
R1 and C/1965 S1, and ever since 1972 it has been listed,
as the representative orbit for C/1945 X1, in the Cata-
logue of Cometary Orbits (see Marsden & Williams 2008
for the most recent edition).
More recently, this comet's orbital motion was further
examined by Marsden (1989). He computed four differ-
ent gravitational solutions, A -- D, constraining them to a
prescribed orientation of the line of apsides and allowing
the orbit to depart from a parabola. His solutions C and
D are discussed more extensively in Section 6.1.5
In terms of the orientation of the line of apsides, a
very stable orbital parameter, the representative orbit of
C/1945 X1 is compared in Figure 1 with more reliably de-
termined orbits of seven bright sungrazers. The errors for
six of them are smaller than the size of the symbols. For
C/1945 X1 the error is estimated by comparing the scat-
ter in the apsidal line with that in the angular elements,
combined with the errors of the elements, as published
by Marsden (1967). The estimated error for C/1880 C1
is due largely to the uncertainty in the orbital period, as
investigated by Kreutz (1901). There is a striking dis-
crepancy between C/1945 X1 and the other sungrazers in
Figure 1, exceeding six standard deviations, thus provid-
ing grounds for suspecting that the motion of C/1945 X1
might be -- like the motions of the dwarf sungrazers --
significantly affected by nongravitational forces.
As of now, it has not been demonstrated conclusively
whether there exists a purely gravitational orbital so-
lution for C/1945 X1 that simultaneously (i) fits satis-
factorily at least four of the five available observations;
(ii) complies with the proper orientation of the line of ap-
sides; and (iii) is consistent with a plausible osculating
orbital period. Addressing this issue requires in the first
place that we know the appropriate values of the quan-
5 Marsden (1989) remarked that the December 14 observation
was reconstructed for an arbitrarily chosen observation time from
the residuals in his previous paper (Marsden 1967), because the
original data were unfortunately lost.
4
Sekanina & Kracht
tities under (ii) and (iii), which depend on the perturba-
tions by the planets, Jupiter in particular. We examine
these perturbation effects next.
5.
INDIRECT PERTURBATIONS BY THE PLANETS
Because of the nature of their orbits, the Kreutz comets
cannot experience a close approach to the planets, in-
cluding Jupiter. Nevertheless, the sungrazers' orbits are
subjected to indirect planetary perturbations over the
entire revolution about the Sun and, as a result, show
limited variations.
Addressing this problem in some detail, Marsden
(1967) began his investigation of the effects of indirect
perturbations on a fragment, separating from its sungraz-
ing parent at perihelion, by integrating the fragment's
motion over the orbital period to the time tπ of next
perihelion. First, he considered only Jupiter in a circular
orbit in the plane of the ecliptic. He applied the equa-
tions for the variation of arbitrary constants and found
that in this simplified scenario the fragment's orbital el-
ements -- the argument of perihelion ω, the longitude of
the ascending node Ω, the inclination i, and the perihe-
lion distance q -- at time tπ depend on Jupiter's ecliptical
longitude at this same time, ΛJ(tπ), as follows:
ω(ΛJ) = ω0 + Xω sin(ΛJ +Λ0),
Ω(ΛJ) = Ω0 + XΩ sin(ΛJ +Λ0),
i(ΛJ) = i0 + Xi sin(ΛJ+Λ0),
q(ΛJ) = q0 − Xq sin(ΛJ +Λ0−90◦)
= q0 + Xq cos(ΛJ+Λ0),
(1)
where ω0, . . . , q0 are constant terms, Xω, . . . , Xq are
amplitudes (all taken as positive numbers), and Λ0 is a
constant phase shift. Because of Jupiter's nonzero orbital
eccentricity and the deviation of its orbital plane from the
plane of the ecliptic, and also because of the indirect per-
turbations by the other planets, the quasi-periodic vari-
ations of the sungrazers' orbital elements are more com-
plex. Marsden indicated that when only Jupiter's per-
turbations were accounted for, the amplitudes amounted
to approximately Xω = 1◦.1, XΩ = 1◦.4, Xi = 0◦.3, and
Xq = 0.00039 AU = 0.084 R⊙. Marsden further pointed
out that his numerical integrations, which took into ac-
count the perturbations by the planets Jupiter to Nep-
tune and by Pluto, resulted in expressions similar to
Equation (1), with the amplitudes equal, respectively,
to 1◦.6, 2◦.1, 0◦.4, and 0.00046 AU (= 0.099 R⊙).
Since the longitude and latitude of perihelion, Lπ and
Bπ, are related to the three angular elements by
tan(Lπ−Ω) = tan ω cos i,
sin Bπ = sin ω sin i,
(2)
the variation of the line of apsides is described, to a first
approximation, by:
Lπ(ΛJ) = L0 + XL sin(ΛJ+Λ0),
Bπ(ΛJ) = B0 + XB sin(ΛJ+Λ0),
(3)
where
fected by the indirect perturbations. To get a more pro-
found insight into their influence on the orbital motion
of C/1945 X1, we first integrated the orbit of its major
presumed sibling, C/1882 R1 (see Table 1 of Sekanina &
Chodas 2004),6 back in time to the 1106 perihelion. Next
we varied the eccentricity incrementally and ran a set of
orbits, obtained in this fashion, forward in time. Using
the JPL DE405 ephemeris, we accounted for the pertur-
bations by all eight planets, by Pluto, and by the three
most massive asteroids, as well as for the relativistic ef-
fect. The increments were adjusted stepwise so that the
perihelion times ranged between 1939 August 2 and 1951
August 29 at nearly constant intervals of about 100 days
each; the length of the covered time period slightly ex-
ceeded Jupiter's orbital period. Checks showed that the
relative errors accumulated over an integration period of
∼1600 yr did not exceed ∼10−10 in the comet's position
vector and its velocity vector.
The integrated perturbations of the six orbital parame-
ters, ω, Ω, i, Lπ, Bπ, and q, for this period of sungrazers'
arrival to perihelion are presented in Figure 2 as a func-
tion of Jupiter's longitude ΛJ at the perihelion times, tπ.
The variations are not periodic in that the values of the
elements at ΛJ = 0◦ in 1939 and 1951 differ by 0◦.70 in ω,
by 0◦.59 in Ω, by 0◦.12 in i, by 0.068 R⊙ in q, by 0◦.044
in Lπ, and by 0◦.012 in Bπ. From Figure 2 it follows
that the variations for the last four parameters are rea-
sonably well approximated by the thick sine curve, while
those for ω and Ω are seen to deviate quite significantly,
especially in the range of Jupiter's longitudes from 30◦ to
150◦. For at least these two elements the scenario based
on Jupiter in a circular orbit in the plane of the ecliptic
is clearly inadequate.
The perturbation variations in the perihelion longitude
Lπ are a problem. Equation (3) suggests that the curve
should be in phase with the variations in the argument
of perihelion, the longitude of the ascending node, and
the inclination, but Figure 2 demonstrates that it is not.
Closer inspection shows that the reason for this discrep-
ancy is the fact that the predicted amplitude given by
Equation (3) is very close to zero (or, in fact, slightly
negative) and that the actual variations in Lπ in Figure 2
are determined by the second order terms that have been
neglected in Equation (3). For the perihelion latitude Bπ
this is not the case and its variations are indeed in phase
with ω, Ω, and i.
The numerical integration shows that the net ampli-
tudes of the orbital parameters amount to 1◦.52 in ω,
1◦.86 in Ω, 0◦.39 in i, 0◦.11 in Lπ (which at a latitude of
∼35◦ represents an arc of 0◦.09), 0◦.07 in Bπ, and 0.00051
AU or 0.110 R⊙ in q. These are typically within about
10% of the amplitudes given by Marsden for the four el-
ements from Equation (1) above. The overall amplitude
for the line of apsides is 0◦.11, so that the range, 0◦.22, is
generally consistent with the apsidal difference of 0◦.30
between C/1882 R1 and C/1843 D1 accumulated over a
span of two orbital periods (or about 15 -- 17 centuries),
suggested for the age of the Kreutz system by Sekanina
& Chodas (2004, 2007).
XL = XΩ + (Xω cos i0 − Xi cos ω0 sin B0) sec2B0,
XB = (Xω cot ω0 + Xi cot i0) tan B0.
(4)
Marsden noticed that the line of apsides is scarcely af-
6 This presumption is based both on Marsden's (1967, 1989) and
Sekanina & Chodas' (2004) conclusions and on the results of our
own experimentation with the orbit of C/1945 X1, which showed
that it was much easier to align its apsidal orientation with that of
C/1882 R1 than C/1843 D1, the other potential major sibling.
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
5
INDIRECT PLANETARY PERTURBATIONS
70◦
69◦
ω
68◦
67◦
∼∼
142◦.0
i
141◦.5
+35◦.3
Bπ
+35◦.1
② ②
②
②
②
②
②
②
② ②
② ②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
PERIHELION
ARGUMENT OF
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
② ② ② ② ②
② ② ② ② ②
ASCENDING NODE
LONGITUDE OF
② ② ② ②
INCLINATION
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
②
LONGITUDE OF PERIHELION
② ② ② ② ② ② ② ② ②
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
② ② ② ② ② ② ② ② ② ② ② ②
② ② ② ② ② ② ②
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
② ② ② ② ② ② ② ② ② ② ② ② ② ②
② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ② ②
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
② ② ② ② ② ② ②
rrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrrr
qqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqqq
② ② ② ② ② ② ② ② ② ② ② ② ② ② ②
LATITUDE OF PERIHELION
PERIHELION DISTANCE
∼∼
348◦
347◦
Ω
346◦
345◦
283◦.0
Lπ
282◦.8
1.7
q (R⊙)
1.5
0◦
60◦
300◦
ECLIPTICAL LONGITUDE OF JUPITER
120◦
240◦
180◦
360◦
Figure 2. Plots, against the ecliptical longitude of Jupiter ΛJ at the time of a respective sungrazer's perihelion, of the argument of
perihelion ω, the longitude of the ascending node Ω, the inclination i, the perihelion longitude Lπ and latitude Bπ (all equinox J2000), and
the perihelion distance q of the siblings of comet C/1882 R1 separated at the 1106 perihelion, which returned to perihelion between 1939
August 2 and 1951 May 18. The sungrazer C/1945 X1 was at ΛJ = 195◦.23. The thick sine curves show the simple approximations of the
type expressed by Equation (1), the thinner curves are the much better approximations by the N -harmonic Fourier polynomials (N = 4 for
all but the last curve, for which N = 2; see Table 3). Note that the scales for Lπ and Bπ are four times wider than those for ω, Ω, and i.
MEAN RESIDUALS FROM
LEAST SQUARES FITS TO
FOURIER POLYNOMIALS
·
LONGITUDE OF
ASCENDING NODE
③
·
·
·
③
·········
③ ③ ③
③
···········
·
············
···
③ ③ ③ ③
···········
ARGUMENT OF
·
···
PERIHELION
③
······
③
···········
INCLINATION
③ ③ ③ ③ ③ ③
···········
···········
6
Sekanina & Kracht
Table 2
Cataloged Orbit of C/1945 X1 and Orbit of the Hybrid
(Equinox J2000).
Orbital element
Perihelion time tπ (1945 TT)
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
Perihelion distance q(cid:26) (R⊙)
(AU)
Orbit eccentricity
Semimajor axis (AU)
Orbital period (yr)
Longitude of perihelion Lπ
Latitude of perihelion Bπ
Osculation epoch (1945 TT)
C/1945 X1
(catalogeda )
Hybridb
(perturbed)
Dec. 27.9652 Dec. 27.9801
72◦.0619
351◦.2006
141◦.8734
0.007516
1.6147
1.0
∞
∞
283◦.57
+35◦.97
(none)
67◦.8518
345◦.3759
141◦.5608
0.007123
1.5302
0.99992634
96.7
951
282◦.84
+35◦.16
Dec. 26.0
a See Marsden & Williams (2008).
bOrbit of C/1882 R1 integrated to perihelion time of C/1945 X1.
σ
±0.3
±0.2
±0◦.1
③
③
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
·
③
·
·
·
③
·
·
·
·
·
·
·
·
③
·
··
·
·
·
③
·
··
·
·
·
We now constrain the orbital properties of C/1945 X1
by its arrival time at perihelion, 1945 December 28.0 TT.
The predicted orientation of the comet's line of apsides
is determined by the ecliptical coordinates Lπ = 282◦.84
and Bπ = +35◦.16 (Equinox J2000), deviating from the
orientation of the apsidal line of C/1882 R1 by 0◦.11.
The longitude of Jupiter at the time of the comet's per-
ihelion is ΛJ = 195◦.23.
In addition, the computations
show that the comet's osculating semimajor axis should
have been 96.7 AU and the orbital period 951 yr, longer
than the actual time span since 1106. This orbital period
is in excellent agreement with the value that Sekanina &
Chodas (2004) obtained for this comet in their hierarchy
model of the Kreutz system (see their Table 12).
To illustrate the influence of the indirect perturbations
specifically on the orbit of C/1945 X1, we compare Mars-
den's (1967) cataloged parabolic orbit (cf. Marsden &
Williams' 2008) with the orbit of what we call a hybrid
-- one that C/1882 R1 would have had, if it arrived at
perihelion at the time C/1945 X1 did. The two orbits
differ by several degrees in the argument of perihelion
and the longitude of the ascending node, and by almost
0.1 R⊙ in the perihelion distance.
Returning now to Figure 2, one could think of the dif-
ferences between the sequence of the points, computed
by numerical integration of the orbits, and the thick sine
curves that are intended to fit them, as deformations and,
accordingly, employ the superposition principle to miti-
gate the discrepancies. The effects of the perturbations,
even though not periodic on a scale of Jupiter's orbital
period, can nonetheless be expressed as a combination
of periodic variations in terms of an N -harmonic Fourier
polynomial. Figure 2 shows the results of this fitting,
with N not exceeding 4, as the lighter curves; obviously,
the improvement over a simple sine curve is considerable.
Thus, calling ℜ any of the six parameters in Equations
(1) and (3), we can write it in the form:
ℜ(ΛJ) =ℜ0 +
=ℜ0 +
N
Xk=1
Xk=1
N
[aℜ,k sin(kΛJ) + bℜ,k cos(kΛJ)]
Xℜ,k sin(kΛJ+Λℜ,k−1),
(5)
1
2
3
4
5
6
7
8
HARMONIC N OF FOURIER POLYNOMIAL
Figure 3. Mean residuals σ of the Fourier polynomials, fitted to
the variations in the longitude of the ascending node, the argument
of perihelion, and the inclination during Jupiter's orbital period
1939 -- 1951, as a function of the polynomials' harmonic N . The
minimum residual for all three orbital elements is reached at N = 4.
where
Xℜ,k =qa2
ℜ,k + b2
Λℜ,k−1 = arctan(cid:18)bℜ,k
aℜ,k(cid:19) .
ℜ,k,
(6)
It is apparent that for N = 1, Equation (5) emulates
Equations (1) and (3), with Λℜ,0 being always equal
to Λ0 for ℜ = (ω, Ω, i, Lπ, Bπ) and Λℜ,0 = Λ0 +90◦ for
ℜ = q, and with Xℜ,1 = Xℜ for any orbital parameter.
The coefficients aℜ,k and bℜ,k are computed by least
squares from the values of ℜ(ΛJ) derived from the or-
bit integration runs as a function of Jupiter's longitude
ΛJ(tπ). The appropriate number of terms N = Nmin > 1
is determined by a minimum mean residual. As the
number of Fourier coefficients in the equations of con-
dition increases with N , the number of the degrees of
freedom drops, thus causing the mean residual σ to in-
crease. A byproduct is an increasing uncertainty of
the Fourier coefficients in the (meaningless) polynomi-
als with N > Nmin. As examples, the plots of the mean
residuals for ω, Ω, and i vs N are displayed in Figure 3.
The Fourier solutions for the six orbital parameters in
Figure 2 are listed in Table 3. For each parameter the
first line provides the results of the approximate solution,
fit A, by applying Equation (1) or (3). The second and
following lines refer to the fitted Fourier polynomial, FN .
The mean residual σ is always followed by the constant
term ℜ0 and by information on each harmonic k in the
case of the N -harmonic Fourier polynomial: the harmonic
amplitude Xℜ,k and phase Λℜ,k−1.
We note that the amplitudes of all harmonics are de-
termined with an error not exceeding ±35%, but mostly
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
7
Fourier Polynomials Fitting the Dependence of Indirect Planetary Perturbations of Orbital Elements of
Kreutz Sungrazers in 1939 -- 1951 on Ecliptical Longitude of Jupiter.
Table 3
Orbital element
Fita
Mean
residual
Constant
termb
Harmonic
k = 1, . . . ,N
Harmonic
amplitude
Harmonic
phase
Argument of perihelion, ω
Longitude of ascending node, Ω
±0.338
±0.113
346.214 ± 0.072
346.216 ± 0.024
±0◦.269
±0.084
68◦.454 ± 0◦.057
68.455 ± 0.018
±0.061
±0.023
141.709 ± 0.013
141.710 ± 0.005
A
F4
A
F4
A
F4
1◦.397 ± 0◦.083
1.386 ± 0.026
0.272 ± 0.026
0.183 ± 0.025
0.105 ± 0.025
1.731 ± 0.104
1.718 ± 0.035
0.338 ± 0.035
0.229 ± 0.034
0.134 ± 0.035
0.364 ± 0.019
0.361 ± 0.007
0.059 ± 0.007
0.042 ± 0.007
0.024 ± 0.007
356◦.2 ± 3◦.3
355.9 ± 1.0
202.1 ± 5.3
57.3 ± 8.0
319 ± 14
359.1 ± 3.3
358.8 ± 1.1
205.8 ± 5.8
59.0 ± 8.7
323 ± 15
357.0 ± 2.8
356.7 ± 1.1
198.6 ± 6.7
52.0 ± 9.6
320 ± 17
104.6 ± 3.9
104.1 ± 1.7
299.2 ± 6.9
143 ± 20
29 ± 18
354.4 ± 4.6
353.8 ± 0.9
194.6 ± 2.7
66.3 ± 4.7
321.4 ± 9.3
84.0 ± 2.9
83.2 ± 2.6
325 ± 13
a A = approximation by Equation (1) or (3), based on the assumption of a single planet in circular orbit in the plane of the ecliptic;
FN = N -harmonic Fourier polynomial that provides the best fit (minimum mean residual).
b Equinox J2000.
0.0968 ± 0.0060
0.0906 ± 0.0026
0.0222 ± 0.0026
0.0077 ± 0.0027
0.0082 ± 0.0027
0.0616 ± 0.0052
0.0608 ± 0.0009
0.0185 ± 0.0009
0.0111 ± 0.0009
0.0055 ± 0.0009
0.0987 ± 0.0048
0.0905 ± 0.0039
0.0173 ± 0.0040
+35.1860 ± 0.0036
+35.1859 ± 0.0006
. . .
1
2
3
4
. . .
1
2
3
4
. . .
1
2
3
4
. . .
1
2
3
4
. . .
1
2
3
4
. . .
1
2
282.9179 ± 0.0043
282.9183 ± 0.0019
1.6200 ± 0.0034
1.6183 ± 0.0028
A
F4
A
F4
A
F2
±0.0202
±0.0087
±0.0169
±0.0030
±0.0156
±0.0132
Orbit inclination, i
Longitude of perihelion, Lπ
Latitude of perihelion, Bπ
Perihelion distance, q (R⊙)
less than ±25%. With one borderline exception the
amplitudes decrease with increasing harmonic, as ex-
pected. The amplitude of the first harmonic is always
only slightly smaller than the net amplitude derived di-
rectly from the results of numerical integration of the or-
bit. The phase angles of all harmonics for the argument
of perihelion, the longitude of the ascending node, the in-
clination, and the latitude of perihelion are remarkably
close to one another. The first harmonic for the perihe-
lion distance is shifted, as expected, by almost exactly
90◦ relative to those of ω, Ω, i, and Bπ. Only the lon-
gitude of perihelion Lπ appears to be out of phase with
any of the other orbital parameters.
The least-squares procedure also allows us to generalize
Equation (5) thus:
ℜ(ΛJ) = ℜ0 +
N
Xk=1
Xℜ,k sin(ckΛJ+Λℜ,k−1),
(7)
where ck's are any positive numbers, not restricted to
integers. This generalization offers us an opportunity to
relate the coefficients ck (k > 1) to the orbital periods of
other planets, Pk, so that ck = PJ/Pk. Without provid-
ing any details, we point out that our search for least-
squares solutions of this type proved much less successful
in fitting the quasi-periodic variations in the sungrazers'
orbital elements than did the Fourier polynomials.
6. ORBITAL COMPUTATIONS BASED ON THE BOYDEN
OBSERVATIONS FROM DECEMBER 11 -- 15
If comet C/1945 X1 is closely related to C/1882 R1,
the hybrid's orbit in Table 3 suggests that the line of
apsides be described by Lπ = 282◦.84 and Bπ = +35◦.16
(Equinox J2000), as plotted in Figure 1 (marked Pert.),
and that the osculating semimajor axis at perihelion be
equal to a = 96.7 AU. Table 3 also shows that the oscu-
lating elements should differ from those of C/1882 R1 by
∼2◦ in ω and Ω, by almost 0◦.5 in i, and by more than
0.1 R⊙ in q. Below we investigate a range of orbital solu-
tions, gravitational and nongravitational, offered by the
observations made at Boyden on December 11 -- 15.
As measured and reduced by Mowbray and listed for
the equinox B1950 by Marsden (1989), the five astro-
metric positions are for the equinox J2000 presented in
Table 4. Because on the first two nights the comet was
imaged with the Cooke lens, while on the three remain-
ing nights with the Metcalf refractor, we assign the posi-
tional data weights that are inversely proportional to the
plate scales of the two instruments. Comparison of the
observation times in column 2 of Table 4 with those in
column 2 of Table 1 shows that the first ones are system-
atically too early by 0.00021 to 0.00022 day, or 18 -- 19 s.
The nature of the difference is unknown; it cannot come
from a longitude discrepancy, because it is equivalent to
8
Sekanina & Kracht
Table 4
Measured and Reduced Boyden Observations of Comet C/1945 X1 (Equinox J2000).
Ref.
No.
Observation time
1945 (UT)
1
2
3
4
5
December 11.04687
12.04691
13.07189
14.07000a
15.06885
R.A.
Dec.
Assigned
weight
Instrument
employed
Plate
number
h
m
s
24
13
.42
15
15 43 21.25
16 11 04.21
16 34 56.77
16 55 41.67
◦
′
′′
43
23
−65
.4
−64 09 13.1
−62 29 32.5
−60 33 16.5
−58 22 47.2
1.0
1.0
3.6
3.6
3.6
Cooke
Cooke
Metcalf
Metcalf
Metcalf
AM 25201
AM 25206
MF 34987
MF 34991
MF 34994
From plate center
distance
P.A.
11◦.94
2.89
1.07
1.16
0.95
124◦.4
68.0
300.5
302.6
285.7
a Position reconstructed by Marsden for an arbitrarily chosen time on this date (see footnote in Section 4).
a distance of as much as ∼500 meters. The UT times in
Table 1, determined from the sidereal times of the be-
ginning and end of the exposures, are in fact UT1 times
and have independently been checked; they never left an
unexplained difference of more than ∼1 s. Nevertheless,
for the sake of comparison with Marsden's (1967, 1989)
results, we retain for this exercise his UT times.
6.1. Gravitational Solutions
Since the observed orbital arc is too short to reliably
compute the eccentricity, it needs to be determined from
other considerations. As already mentioned above, the
constraint is here provided by adopting the semimajor
axis of the hybrid's osculating orbit, 96.7 AU at the epoch
of 1945 December 26.0 TT, for all solutions that follow.
We begin with a gravitational orbit to fit all five astro-
metric positions in Table 4.
Deriving orbits B and C and comparing them with or-
bit A, Marsden (1989) demonstrated that elliptical solu-
tions fitted the five observations of C/1945 X1 much bet-
ter than a parabolic approximation. Yet, neither orbit,
B or C, provides a satisfactory fit. One possible contrib-
utor to this problem could be the constraint introduced
to satisfy the prescribed direction of the line of apsides
(for which Marsden took Lπ = 282◦.7, Bπ = +35◦.2 af-
ter conversion to Equinox J2000), the same issue that
we encountered with the dwarf sungrazers of the Kreutz
system (Paper 1; see also Section 1).
Comparison of orbit C, the more realistic of the two
sets in terms of the comet's orbital period, with our opti-
mum gravitational fit to the five observations is presented
in Table 5. Marsden's (1989) residuals from orbit C in
this table have a tendency to become more positive to-
ward both ends of the observed arc in right ascension and
to get more negative with time in declination. We tried
to emulate the residuals from orbit C, employing the set
of truncated elements in Marsden's (1989) Table V, but
succeeded to achieve this within about 4′′ only in decli-
nation. Although we allowed for the rounding off of the
perihelion time, the residuals in right ascension came out
systematically more negative by about 12′′ and all except
the last one were negative. This effect is due apparently
to the elements' truncation.
The residuals from our Best Fit solution, although
clearly better than those from orbit C, are not quite
satisfactory either, suggesting that the first observation
may be inferior. That should not be surprising, consid-
ering that the comet was rather far from the plate center
on this discovery exposure (Section 3). Unfortunately,
the line of apsides deviates from the proper direction by
0◦.571, which is unacceptably large. Although the in-
dividual angular elements of the Best Fit solution are
burdened with errors of up to several tenths of a degree,
the correlation among them draws the uncertainty in the
orientation of the line of apsides down to only hundredths
of a degree.
Next, we compare Marsden's (1989) orbit D with the
hybrid's orbit from Table 2. As with orbit C, we first
tested whether the low-precision orbit D from Marsden's
Table V is able to reproduce the residuals in his Table VI.
Since this exercise did not meet with success, we decided
to reconstruct the high-precision version of orbit D, not
published by Marsden, by subtracting the residuals in
his Table VI from the observed astrometric positions, re-
taining his constraint on the reciprocal semimajor axis
(0.01016 AU−1). This approach was successful; the re-
sulting orbit D is shown in Table 6 to be exceedingly
similar to the hybrid's orbit, the two sets of elements dif-
fering mostly in the fourth or higher significant digit. In
particular, we note that the lines of apsides agree to bet-
Table 5
Marsden's (1989) Orbit C and Our Optimum Gravitational
Solution to Fit All Five Boyden Observations of
C/1945 X1 (Equinox J2000).
Orbital element
Orbit C
Best Fit
Perihelion time tπ (1945 TT)
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
Perihelion distance q(cid:26) (R⊙)
(AU)
Orbit eccentricity
Semimajor axis (AU)
Orbital period (yr)
Longitude of perihelion Lπ
Latitude of perihelion Bπ
Osculation epoch 1945 (TT)
Mean residuala
Dec. 27.976 Dec. 27.982
68◦.10
345◦.54
141◦.60
0.00699
1.502
70◦.99
349◦.50
141◦.91
0.007244
1.556
0.99993010
0.99992506
100
1000
282◦.70
+35◦.20
(none)
±11′′.95
96.67
950.5
283◦.14
+35◦.68
Dec. 11.0
±6′′.17
Distribution of Residualsa, O−C
Time of
observation
1945 (UT)
Orbit C
Best Fit
R.A.
Dec.
R.A.
Dec.
Dec. 11.04687
12.04691
13.07189
14.07000
15.06885
+8′′.0
+2.5
−5.6
+8.0
−3.2
aFrom unweighted observations for orbit C, weighted for Best Fit.
+5′′.6 +22′′.4
+5.3
+6.1
−11.8
−14.8
−1.9
+10.3
−10.9
+15.6
+13′′.2
+6.1
−6.6
+7.8
−2.7
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
9
Marsden's (1989) Orbit D and Hybrid's Orbit (Equinox J2000).
Table 6
Orbital element
Orbit D
Hybrid's orbit
Perihelion time tπ (1945 TT)
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
Perihelion distance q(cid:26) (R⊙)
(AU)
Orbit eccentricity
Semimajor axis (AU)
Orbital period (yr)
Longitude of perihelionb, Lπ
Latitude of perihelionb, Bπ
Osculation epoch 1945 (TT)
Mean residualc
Dec. 27.9798a Dec. 27.9801
67◦.8356
345◦.3571
141◦.5553
0.007129
1.5315
67◦.8514
345◦.3753
141◦.5607
0.007123
1.5302
0.99992756
0.99992632
98.42
976.4
282◦.84
+35◦.16
Dec. 11.0
(cid:26) ±7′′.63 (M)
±8′′.17 (SK)
96.67
950.5
282◦.84
+35◦.16
Dec. 11.0
±13′′.75
Distribution of Residualsc, O−C
Time of
observation
1945 (UT)
Orbit D (M) Orbit D (SK) Hybrid's orbit
R.A. Dec.
R.A. Dec.
R.A. Dec.
Dec. 11.04687 +10′′.8 +5′′.1 +12′′.3 +6′′.6 +26′′.4 −6′′.3
−1.1 −1.1 +13.7 −12.1
+1.2 −18.7
−14.0 −9.6
+5.3 +7.6 +20.4
0.0
+3.6 +4.5 +18.4 −1.8
−2.5 −2.8
12.04691
13.07189 −15.2 −11.4
14.07000
+4.3 +5.6
+2.8 +2.5
15.06885
aTo fit weighted observations, tπ = Dec. 27.9800.
bThese ecliptical oordinates define the reference line of apsides.
cEntry (M) and columns 2 -- 3 from unweighted observations; entry
(SK) and columns 4 -- 5 from weighted observations; residuals from
hybrid's orbit are from weighted observations.
ter than 0◦.01. This correspondence is not fortuitous, as
both sets of elements represent perturbed versions of the
orbits of two comets observed 83 years apart, whose mo-
tions in space were virtually identical, as already pointed
out by Marsden (1967): C/1882 R1, used by us to derive
the hybrid's orbit; and C/1965 S1, employed by Marsden
(1989) to derive orbit D.
Table 6 also lists three sets of residuals: those left by
orbit D from the unweighted observations, in columns 2
and 3 [taken from Marsden's (1989) Table VI and marked
M]; those left by orbit D from the weighted observations,
as derived by us7 (marked SK), in columns 4 and 5; and
those left by the hybrid's orbit from the weighted obser-
vations, in columns 6 and 7.
Even though the residuals in Table 6 are fairly large,
the ones left by orbit D are much better than the strongly
systematic ones left by the hybrid's orbit. From this com-
parison as well as from the difference in the mean residual
we infer that the orbital evolution of comet C/1945 X1,
as a fragment of its common parent with C/1882 R1 and
C/1965 S1, was apparently more similar to the orbital
evolution of the latter than the former. This is an im-
portant though tentative conclusion, consistent with the
fragmentation hierarchy of the Kreutz system proposed
by Sekanina & Chodas (2004). According to their evolu-
tionary model, C/1882 R1, C/1965 S1, and the precur-
sor to C/1945 X1 separated from their common parent,
possibly X/1106 C1, at the same time, some 18 days
7 This fit required that the perihelion time be increased by 0.0002
day relative to the tabulated value; otherwise the residuals would
be slightly greater.
after perihelion. Relative to C/1882 R1, the precursor
moved in the same direction as, and with a velocity only
about 20% lower than, C/1965 S1 (see also Sekanina &
Chodas 2002). The subsequent separation of C/1945 X1
from its precursor around 1700 AD notwithstanding, the
comet's orbit in 1945 should indeed resemble the orbit of
C/1965 S1 to a greater degree than that of C/1882 R1.
Since orbit D and the hybrid's orbit are so very similar,
yet the residuals they leave substantially differ, we tested
whether this effect is due to the minor differences in the
angular elements or the orbital dimensions. We replaced
the semimajor axis 98.4 AU with 96.7 AU and noted that
the residuals from orbit D in Table 6 did not change; the
new orbit, D′, is now: tπ = 1945 December 27.9803 TT,
q = 0.007126 AU = 1.5309 R⊙, e = 0.99992630, and for
the equinox of J2000, ω = 67◦.8377, Ω = 345◦.3558, and
i = 141◦.5570. The residuals in Table 6 reflect thus entire-
ly the differences of up to 0◦.02 in the angular elements.
Before turning to nongravitational solutions, we take
notice of a possibility that one of the five astrometric
observations is inferior and should be discarded before
computing any orbit. Accordingly, we are now going to
search for solutions that could fit four observations. Such
solutions will be referred to as G{ℑ}, where G stands for
gravitational and {ℑ} is a progression of reference num-
bers of the observations that such solutions are based on;
these numbers are listed in column 1 of Table 4. The Best
Fit solution presented in Table 5 can now be referred to
as G{1, 2, 3, 4, 5}. As already mentioned above, the dis-
covery position may be inferior, so that one of the tested
four-observation solutions is G{2, 3, 4, 5}. The residuals
from orbit D in Table 6 suggest that the third observa-
tion may be even worse than the first, raising interest in
the solution G{1, 2, 4, 5}. Finally, the derivation of the
cataloged parabolic orbit implied that the fourth obser-
vation failed to fit that solution (Marsden 1967), hence a
need to test G{1, 2, 3, 5}. The results of all three of these
solutions are presented in Table 7. No orbits are being
computed based on three observations only, as these are
regarded for our purposes as meaningless.
Since a semimajor axis a between 96.7 AU and 98.4 AU
was shown to make no difference, the four-observation
solutions below are subjected to the same constraint as
the Best Fit solution in Table 5, namely, a = 96.7 AU
at an osculation epoch of 1945 December 26.0 TT. The
solutions listed in Table 7 differ from one another consid-
erably. The orbits G{2, 3, 4, 5} and G{1, 2, 3, 5} are both
unacceptable on account of their large offsets from the
reference apsidal direction (given by the hybrid's orbit
and Marsden's orbit D). The set G{2, 3, 4, 5} is in addi-
tion handicapped by the very large residuals left by the
December 12 observation. Even cursory inspection shows
that the orbit G{1, 2, 4, 5} is by far the most promising of
the three, implying that it is the December 13 observa-
tion that is inferior. The offset from the reference line of
apsides, slightly less than 0◦.1, is reasonably low though
not completely satisfactory.
It is somewhat surprising
that the apparently inferior astrometric position is one of
those taken with the Metcalf refractor. We would expect
that the discarded data point should be one of the first
two observations, made with the Cooke camera. Under
the circumstances, the solution G{1, 2, 4, 5} can at best
be regarded as marginally acceptable to approximate the
orbital motion of C/1945 X1.
10
Sekanina & Kracht
Table 7
Comparison of Three Gravitational Solutions Based on
Four Observations of C/1945 X1 (Equinox J2000).
Orbital element
G{2, 3, 4, 5} G{1, 2, 4, 5} G{1, 2, 3, 5}
Perihelion time tπ (1945 TT)
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
Perihelion distance q(cid:26) (R⊙)
(AU)
Orbit eccentricity
Longitude of perihelion Lπ
Latitude of perihelion Bπ
Reference apsidal line's offseta
Osculation epoch 1945 (TT)
Mean residual (weighted)
Dec. 27.986 Dec. 27.979 Dec. 27.983
65◦.98
342◦.84
141◦.37
0.007004
1.505
67◦.43
344◦.82
141◦.50
0.007120
1.530
74◦.00
353◦.52
142◦.07
0.007415
1.593
0.99992755
0.99992635
0.99992329
282◦.54
+34◦.77
0◦.462
282◦.80
+35◦.09
0◦.078
283◦.49
+36◦.22
1◦.184
Dec. 11.0
Dec. 11.0
Dec. 11.0
±5′′.11
±1′′.79
±1′′.06
Distribution of Residualsb, O− C
Time of
observation
1945 (UT)
Dec. 11.04687
12.04691
13.07189
14.07000
15.06885
G{2, 3, 4, 5}
G{1, 2, 4, 5}
G{1, 2, 3, 5}
R.A.
Dec.
R.A.
Dec.
R.A.
Dec.
. . . . . .
. . . . . .
+25′′.2 +12′′.0
−4.3
−5.1
+5.5
+4.3
−1.1
−2.3
+5′′.3 +1′′.1
−7.4
−8.2
. . . . . .
. . . . . .
+1.3
+0.1
−0.8
+0.1
−3′′.3
+4.9
+0.1
. . . . . .
−0.2
+3′′.3
+2.7
−0.8
. . . . . .
+0.3
aReference line of apsides is defined by the hybrid's Lπ and Bπ.
bFrom weighted observations.
In any case, it is worth examining whether any mean-
ingful refinement of the orbit, an apsidal line in particu-
lar, can be achieved by incorporating a nongravitational
acceleration, which, if comparable to those affecting the
motions of the dwarf sungrazers, might be detectable
over a span of four days. Besides, such alternative so-
lutions should prove beneficial to estimating a range of
uncertainties in the comet's orbital motion (Section 7).
6.2. Standard Nongravitational Solutions
In order to find out whether there is evidence for non-
gravitational effects in the motion of comet C/1945 X1
in the meager set of astrometric data available, we be-
gin with a standard formalism of Marsden et al. (1973),
based on a water-ice sublimation model. The erosion-
driven nongravitational accelerations in the three cardi-
nal directions, i.e., in the radial (away from the Sun),
R, transverse, T, and normal, N, directions of a right-
handed RTN coordinate system that is referred to the
comet's orbital plane, are in this formalism expressed by
aR(r)
aT(r)
aN(r)
=
A1
A2
A3
· gstd(r),
(8)
where gstd(r) is the standard law employed by Marsden et
al. (1973), approximating the sublimation rate of water
ice from an isothermal spherical nucleus and normalized
so that gstd(1 AU) = 1,
gstd(r) = α(cid:18) r
r0(cid:19)−m(cid:20)1+(cid:18) r
r0(cid:19)n(cid:21)
−k
.
(9)
Here m = 2.15, n = 5.093, nk = 23.5, a scaling distance
r0 = 2.808 AU, and a normalization constant α = 0.1113.
The parameters A1, A2, A3 are, respectively, the radial,
transverse, and normal components of the nongravita-
tional acceleration at 1 AU from the Sun; their units
usually employed in orbital studies are 10−8 AU day−2.
Since the vector of the acceleration due to the generally
sunward-directed sublimation points away from the Sun,
physically meaningful values of A1 should always be pos-
itive. The law (9) has over the past 40 years served ad-
mirably in countless orbit-determination applications to
comets with perihelia typically several tenths of AU.
The numerical procedure that we apply next follows
the method described in Paper 1. Briefly, since the di-
rection of the line of apsides was found to be a function of
the nongravitational acceleration, a minimum offset from
the reference, or nominal, apsidal line (defined here by
Lπ and Bπ for the hybrid's orbit or orbit D; see Table 6)
serves as the constraint that determines the nongravita-
tional acceleration's most probable magnitude.
Similarly to our notation for the gravitational runs,
we refer to these nongravitational solutions by N (X)
std{ℑ},
where {ℑ} is again a progression of reference numbers of
the employed observations from Table 4, while N (X)
std de-
notes, on the one hand, Marsden et al.'s (1973) formalism
of accounting for the nongravitational effect, with the
standard law (9) describing the variations with heliocen-
tric distance; and, on the other hand, one of the two ver-
sions applied: either solving for the radial component of
the acceleration with the parameter A1 when X = R; or
for the normal component with A3 when X = N. When-
ever we attempted to solve for both components, the run
aborted. We recall that the radial component (A1 > 0)
dominates the magnitude of the nongravitational accel-
eration in the motions of the cataloged comets in nearly-
parabolic orbits (Marsden & Williams 2008), while the
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
11
Table 8
Comparison of the Nongravitational Parameters
and Test Results for Eight Orbital Solutions
Based on Marsden et al.'s formalism.
Nongravitational
solution
Parametera
A1 or A3
N (R)
N (R)
N (R)
N (R)
N (N)
N (N)
N (N)
N (N)
std {1, 2, 3, 4, 5}
std {2, 3, 4, 5}
std {1, 2, 4, 5}
std {1, 2, 3, 5}
std {1, 2, 3, 4, 5}
std {2, 3, 4, 5}
std {1, 2, 4, 5}
std {1, 2, 3, 5}
−0.82 ± 0.17
+0.67 ± 0.18
+0.113 ± 0.019
−1.70 ± 0.40
−1.87 ± 0.23
+1.54 ± 0.30
+0.234 ± 0.018
−4.18 ± 0.65
Apsidal-
lineb
offset
Mean
residual
(weighted)
0◦.116
±6′′.42
0.118
0.013
0.270
0.073
0.085
0.006
0.212
±5.05
±1.73
±3.03
±6.68
±4.88
±1.72
±2.61
a In units of 10−5AU day−2.
bOffset from direction of the reference line of apsides (Table 6).
normal component (A3) was found to contribute signifi-
cantly to the magnitude of the acceleration that affects
the motions of the dwarf Kreutz sungrazers (Paper 1).
std {2, 3, 4, 5}, N (R)
std {1, 2, 3, 4, 5}, N (R)
Table 8 summarizes the most important results from
the runs based on Marsden et al.'s (1973) standard for-
malism. Altogether we computed eight such solutions for
four different observational sets {ℑ}, the same ones as be-
fore: N (R)
std {1, 2, 4, 5},
N (R)
std {1, 2, 3, 5}; and similarly with N instead of R. Both
the R and the N versions of the N (X)
std {1, 2, 3, 4, 5} and
N (X)
std {1, 2, 3, 5} runs are inferior to their respective grav-
itational solutions (Tables 5 and 7) in terms of the mean
residual: ±6′′.42 (R) and ±6′′.68 (N) against ±6′′.17 in
the case of the {1, 2, 3, 4, 5} solution; and ±3′′.03 (R) and
±2′′.61 (N) against ±1′′.06 in the case of the {1,2,3,5}
solution. In addition, for these two sets of solutions the
R-version parameters A1 come out to be negative and
therefore physically meaningless.
Even though the set {2, 3, 4, 5} avoids the misfortunes
of the sets {1, 2, 3, 4, 5} and {1, 2, 3, 5}, its offsets from
the reference line of apsides fail to improve over the
corresponding offsets from {1, 2, 3, 4, 5}; in addition, the
residuals left by the December 12 observation are com-
parable to those from the equivalent gravitational solu-
tion (Table 7) and entirely unacceptable. In summary,
the standard nongravitational solutions based on the set
{1, 2, 4, 5} are by far the best, in terms of both the offset
from the reference apsidal line and the mean residual.
Also, both A1 and A3 of the {1, 2, 4, 5} solutions come
out in Table 8 to be almost or just about one order of
magnitude smaller than those of the poor solutions.
As the final comments we note that (i) relatively to
the gravitational fit in Table 7, both nongravitational
solutions based on the set {1, 2, 4, 5} improve the mean
residual only marginally (from ±1′′.79 to ±1′′.72/1′′.73),
but reduce the offset from the reference apsidal line con-
siderably (from 0◦.078 to 0◦.013/0◦.006); (ii) both A1 and
A3 appear to be fairly well determined (with the relative
errors of ±8 -- 17%) and on the same order of magnitude,
∼10−6 AU day−2, as the parameters of the normal com-
ponent of the nongravitational acceleration of the dwarf
sungrazers C/2009 L5 and C/2006 J9 in Table 4 of Pa-
per 1; and (iii) these parameters are one order of mag-
nitude greater than the peak values of A1 for the cata-
loged comets in nearly-parabolic orbits and at least two
orders of magnitude greater than the typical values of
A1 for such comets (Marsden & Williams 2008). In sum-
mary, there is some -- though, due to the limited data,
not overwhelming -- evidence that in terms of the non-
gravitational effects in its orbital motion, C/1945 X1 may
share the properties of some dwarf Kreutz sungrazers.
6.3. Modified Nongravitational Solutions
The EXORB7 orbit-determination code employed by
the second author allows one to vary arbitrarily all five
parameters of the nongravitational law (9) -- the expo-
nents m, n, k, the scaling distance r0, and the normal-
ization constant α. This option facilitates a more robust
orbital experimentation, when the standard nongravita-
tional model of Marsden et al. (1973) fails to provide sat-
isfactory results. In Paper 1 we gained some experience
with what we hereafter refer to as a modified nongravi-
tational law gmod(r; r0), which is given by Equation (9),
but while the three exponents remain unchanged from
the standard law, the scaling distance varies broadly, al-
ways entailing a change in α as well.
The physics behind this sublimation-law experimenta-
tion involves a parallelism between the scaling distance
r0 and the snow line, a boundary of the zone in vacuum
(or near vacuum), beyond which it is cold enough for a
volatile substance to exist only in the solid phase. Cal-
ibrated by water ice, for which in the isothermal model
r0 ≃ 2.8 AU, this distance depends in the first approxi-
mation on the effective latent heat of sublimation, L (in
Table 9
Orbital Solutions N (R)
std {1, 2, 4, 5} and N (N)
(Equinox J2000).
std {1, 2, 4, 5}
Orbital element
N (R)
std {1, 2, 4, 5} N (N)
std {1, 2, 4, 5}
Perihelion time tπ (1945 TT) Dec. 27.97928 Dec. 27.97938
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
67◦.910
345◦.437
141◦.572
0.0071167
67◦.883
345◦.412
141◦.565
0.0071250
Perihelion distance q(cid:26) (R⊙)
(AU)
Orbit eccentricity
Longitude of perihelion Lπ
Latitude of perihelion Bπ
Reference apsidal-line offset
Parameter A1 (10−5AU day−2)
Parameter A3 (10−5AU day−2)
Osculation epoch 1945 (TT)
Mean residual (weighted)
1.5287
1.5307
0.99992638
0.99992629
282◦.82
+35◦.16
0◦.0133
+0.113a
. . . . . .
Dec. 11.0
±1′′.73
282◦.83
+35◦.16
0◦.0060
. . . . . .
+0.234b
Dec. 11.0
±1′′.72
Distribution of Residuals O−C
Time of
observation
1945 (UT)
Dec. 11.04687
12.04691
14.07000
15.06885
N (R)
std {1, 2, 4, 5}
N (N)
std {1, 2, 4, 5}
R.A.
+4′′.9
−8.2
+0.3
−0.1
Dec.
+2′′.0
−7.3
+1.0
−0.6
R.A.
+4′′.9
−8.0
+0.2
0.0
Dec.
+1′′.6
−7.2
+1.1
−0.7
aWith a mean error of ±0.019 × 10−5AU day−2 (Table 8).
bWith a mean error of ±0.018 × 10−5AU day−2 (Table 8).
12
Sekanina & Kracht
cal mol−1), of the sublimating species:
L (cid:19)2
r0 =(cid:18) const
,
(10)
1
where r0 is expressed in AU and the constant is equal
2 cal mol−1. Marsden et al. (1973) showed
to 19 100 AU
that in a plot of log (normalized sublimation rate) against
log (heliocentric distance) the value of r0 shifts the curve
left or right along the axis of heliocentric distance r.
Hence, by properly choosing r0, the erosion-driven non-
gravitational effects in the orbital motion of any comet
can approximately be expressed by a universal curve of
log (normalized sublimation rate) against log(r/r0).
Highly refractory materials, such as metals or silicates,
can sublimate only very close to the Sun, so that their
snow line and scaling distance are much smaller than
1 AU. For example, for forsterite, the Mg end-member
of the olivine solid solution system, the latent heat of
sublimation is 130 000 cal mol−1 (Nagahara et al. 1994),
so that its r0 ≃ 0.02 AU.
In limiting cases the empirical equation (9) offers the
following expressions for variations in the sublimation
rates (equally applying to the modified and standard
laws):
lim
x→0
lim
x→∞
gmod(r; r0)∼ r−2.15,
gmod(r; r0)∼ r−25.65.
(11)
where x = r/r0. The first limit approaches an extreme
scenario in which all incident solar energy is spent on sub-
limation, while the second limit crudely approximates the
other extreme, when the energy is spent entirely on heat-
ing the object (and increasing its thermal reradiation).
The actual limiting expressions are r−2 for r/r0 → 0 and
(12)
RT (cid:19) for r/r0 → ∞,
γmod(r; r0) ∼ exp(cid:18)−
L
where the sublimation heat L is related to r0 by Equation
(10), R is the gas constant, and T is the temperature at
heliocentric distance r. The differences relative to (11)
are due to the approximate nature of the law (9), which
is responsible for a peculiar feature when applied to the
orbital motion of C/1945 X1 (see below). To illustrate
it, we begin by introducing a local slope ζ of a modified
nongravitational law between r and r + ∆r; similarly to
(11), it is
lim
∆r→0
gmod(r; r0) ∼ r−ζ ,
(13)
where
∂ ln gmod(r; r0)
nk
= m +
∂ ln(r/r0)
ζ(r; r0) = −
1 + (r/r0)−n . (14)
The limits on ζ are those given in (11). Since r0 is nearly
3 AU in the standard law, the contribution to the inte-
grated nongravitational effect from r≫ r0 is always neg-
ligible and the exact shape of the standard nongravi-
tational law at heliocentric distances r≫ r0 is unimpor-
tant. However, when r0 in a modified law is considerably
smaller than that of the standard law, the power-law ap-
proximation to the exponential law for the sublimation
·······························································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································································
105
1
10−5
10−10
10−15
10−20
10−25
10−30
M
O
D
I
F
I
E
D
N
O
N
G
R
A
V
I
T
A
T
I
O
N
A
L
L
A
W
E
F
F
E
C
T
✄
r∗/r0
✄
✄
✄
✄
✄
gmod(r; r0)
PPPPPPP✐
∗
··············
·····················
·····················
·····················
·····················
·····················
✄✄✎
·····················
·····················
·····················
·····················
·····················
·····················
·····················
·····················
·····················
·····················
·····················
················································································································································································································································································································································································································································································································································································································································································
✏✏✏✏✏✏✶
γmod(r; r0)
MODIFIED
NONGRAVITATIONAL
LAW gmod(r; r0) AND
SUBLIMATION RATES
AT DISTANCES r > r∗
0.1
0.2
HELIOCENTRIC DISTANCE RATIO r/r0
0.4
0.7
1
2
4
7
10
Figure 4. The nongravitational effects in the orbital motion of
a comet derived from the modified law gmod(r; r0) applicable to
an arbitrary scaling distance r0. At heliocentric distances r > r∗
the law γmod(r; r0) matches the exponential variations in the sub-
limation rate better than gmod(r; r0). The laws gmod(r; r0) and
γmod(r; r0) have equal logarithmic slopes at r∗ . The slope ζ of
gmod(r; r0) at r > r∗ is practically constant.
rate beyond r0 could play a major role for comets whose
perihelion distances q are much greater than r0.
To find out at what heliocentric distance r∗ does a
modified law begin to deviate from the sublimation law
γmod(r; r0), we first define the local slope ξ of γmod(r; r0)
similarly to (14):
ξ(r; r0) = −
∂ ln γmod(r; r0)
∂ ln(r/r0)
L√r
2RT0
L√r0
2RT0r r
r0
,
=
=
(15)
where T (r) = T0/√r at the heliocentric distances of ex-
tremely low sublimation rates; for the isothermal model,
T0 = 278.3 K. Since, from Equation (10), L√r0 is a con-
stant and L√r0/RT0 = 34.54, the point of deviation of
the modified law gmod(r; r0) from the exponential sub-
limation law, given by the condition of equal slopes,
ξ(r∗ ; r0) = ζ(r∗ ; r0), determines r∗ by comparing (14)
with (15). After inserting the numerical values,
r∗ = 2.12 r0.
(16)
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
13
Less than a millionth part of the Sun's incident energy
at the distance r∗ is spent on the sublimation.
Requiring that gmod(r∗ ; r0) = γmod(r∗ ; r0) and writ-
ing the exponential sublimation law in the form
γmod(r; r0) = β exp(cid:20)34.54(cid:18)1 −r r
r0 (cid:19)(cid:21) ,
we have
β = 0.0267α.
(17)
(18)
The relationship between gmod(r; r0) and γmod(r; r0) is
displayed in Figure 4. At present, the γmod(r; r0) law is
not incorporated into the orbit-determination code that
we employ, as in most cases it would make hardly any
difference numerically. However, the approximation of
γmod(r; r0) by gmod(r; r0) at r > r∗ (i) prevents one from
testing nongravitational laws with a variable exponent ζ
at these heliocentric distances and (ii) yields identical
results from all solutions based on the scaling distances
r0 that are by more than 2.12 r0 smaller than the least
heliocentric distance at which the comet is observed, as
seen from Figure 4.
In Paper 1 we employed the modified-law paradigm to
great advantage. We determined for all eight in-depth
examined dwarf sungrazers that the astrometric posi-
tions were fitted better (and the offsets from the ref-
erence line of apsides came out to be smaller) when the
nongravitational acceleration affecting their orbital mo-
tions was assumed to vary with heliocentric distance r
much more steeply than prescribed by the standard law
gstd(r). The orbital motions of six out of the eight cases
were fitted best with r0 < 0.13 AU, which corresponds to
L >∼ 50 000 cal mol−1, and for none of the eight was r0
higher than ∼2.1 AU. The heliocentric distances of the
dwarf Kreutz sungrazers in Paper 1 varied between 0.037
and 0.068 AU (or between 8 and 15 R⊙), so that these ob-
jects were considerably closer to the Sun than C/1945 X1
when under observation from Boyden.
For C/1945 X1, the modified-law nongravitational so-
lutions with r0 ≪ 0.2 AU fit most satisfactorily the pre-
ferred set of four observations, {ℑ0} = {1, 2, 4, 5}. This
match is better than that from the standard-law solu-
tions listed in Table 9. We refer to these modified-law so-
lutions as N (X)
mod[r0]{ℑ0}, where X has the same meaning
as before and r0 is in AU. Because the comet was not ob-
served closer to the Sun than 0.6 AU, all these solutions
are identical, as pointed out more generally under (ii)
below Equation (18). We confirm this result by running
solutions for r0 equal to 0.01, 0.05, and 0.10 AU. We then
select, as examples, the solutions N (R)
mod[0.05]{ℑ0} and
N (N)
mod[0.05]{ℑ0} and present them in Table 10; relative
to the results based on the standard law, the offsets from
the reference line of apsides now dropped, respectively,
from 0◦.013 to 0◦.005 and from 0◦.006 to 0◦.003, while
the weighted mean residuals dropped, respectively, from
±1′′.73 and ±1′′.72 down to ±1′′.66 in both R and N.
The nongravitational parameters A1 and A3 are now
determined with relative errors of only ±5 -- 7%. Their
apparent discrepancy of nearly five orders of magnitude
compared to those of the standard law is due entirely
to the very different steepness of the two laws' slopes.
Since the comet was observed between 0.718 AU from
the Sun (on December 11) and 0.598 AU from the Sun
Table 10
Orbital Solutions N (R)
mod[0.05]{ℑ0} and N (N)
(Equinox J2000).
mod[0.05]{ℑ0}
Orbital element
N (R)
mod[0.05]{ℑ0} N (N)
mod[0.05]{ℑ0}
Perihelion time tπ (1945 TT) Dec. 27.97937 Dec. 27.97940
Argument of perihelion ω
Longitude of ascending node Ω
Orbit inclination i
67◦.869
345◦.397
141◦.561
0.0071279
1.5313
0.99992626
282◦.84
+35◦.16
0◦.0034
. . . . . .
+0.877b
Dec. 11.0
±1′′.66
67◦.876
345◦.403
141◦.562
0.0071256
1.5308
Perihelion distance q(cid:26) (R⊙)
(AU)
0.99992629
Orbit eccentricity
282◦.83
Longitude of perihelion Lπ
+35◦.16
Latitude of perihelion Bπ
0◦.0055
Reference apsidal-line offset
Parameter A1 (10−10AU day−2) +0.307a
Parameter A3 (10−10AU day−2)
. . . . . .
Osculation epoch 1945 (TT)
Mean residual (weighted)
±1′′.66
Dec. 11.0
Distribution of Residuals O−C
Time of
observation
1945 (UT)
Dec. 11.04687
12.04691
14.07000
15.06885
N (R)
mod[0.05]{ℑ0}
N (N)
mod[0.05]{ℑ0}
R.A.
+4′′.7
−8.0
+0.3
−0.1
Dec.
+2′′.0
−6.9
+0.9
−0.5
R.A.
+4′′.8
−7.9
+0.3
0.0
Dec.
+1′′.7
−6.9
+1.0
−0.6
aWith a mean error of ±0.021 × 10−10AU day−2.
bWith a mean error of ±0.046 × 10−10AU day−2.
(on December 15), the magnitudes of the effective non-
gravitational accelerations need to be compared in this
range of heliocentric distances, rather than at 1 AU
from the Sun, to test whether the results from the stan-
dard law and the modified law are generally consistent.
This test shows that the radial components from the
two laws reach the same value, +0.3 × 10−5 AU day−2,
at r = 0.639 AU, while the normal components equally
reach +0.6 × 10−5 AU day−2 at r = 0.648 AU, in either
case well inside the observed range of heliocentric dis-
tances. The comet's favorable comparison with the dwarf
sungrazers, mentioned in Section 6.2, is consistent with
the results from the modified-law solutions: C/1945 X1
appears to be subjected to a nongravitational accelera-
tion that is near the lower end of the range of accelera-
tions typical for the dwarf sungrazers.
In summary, we obtain some evidence suggesting that
(i) the nongravitational solutions fit the observations of
C/1945 X1 better than the gravitational solutions; and
(ii) a law with a very steep slope (of the type ∼r−25) may
be superior to the standard law. Over all, however, the
promising performance of the modified law has not over-
turned our conclusion expressed at the end of Section 6.2
that the limited data sample does not make a conclusion
about the existence of large nongravitational effects in
the motion of C/1945 X1 sufficiently compelling. Every
effort should be expended to improve the comet's orbit
by extending the observed arc with the help of additional
images on Boyden patrol plates.
6.4. Astrometric and Orbital Accuracy
It may seem to be questionable to use the astrometric
data of less than high quality for supporting the type of
14
Sekanina & Kracht
Table 11
Geocentric Ephemeris for Comet C/1945 X1 from 1945 September 12 to 1946 January 30 (Equinox J2000).
From solution N (N)
mod[0.05]{ℑ0}
Date TT
R.A.
Dec.
∆
r
App. magnitude
Differential ephemerides from solutions
Phase
angle PA(RV )
r−4
r−7
N (N)
std {ℑ0}
G{ℑ0}
orbit D ′
1945 Sept. 12
22
Oct. 2
12
22
Nov. 1
11
21
Dec. 1
6
11
16
1946 Jan. 5
10
20
30
◦
′
h
m
39
.0 −10
7
00
7 50.1 −11 48
8 01.6 −14 03
8 13.7 −16 54
8 27.0 −20 36
8 42.3 −25 33
9 02.4 −32 30
9 35.1 −42 48
10 54.6 −57 49
12 34.0 −65 04
15 11.8 −65 27
17 12.4 −56 11
18 32.1 −42 17
19 14.6 −55 37
0 37.9 −69 21
3 32.8 −48 43
2.860
2.592
2.315
2.031
1.744
1.458
1.180
0.922
0.713
0.645
0.617
0.640
0.662
0.566
0.525
0.664
2.466
2.310
2.148
1.980
1.804
1.619
1.423
1.213
0.982
0.855
0.719
0.569
0.434
0.601
0.882
1.124
20◦.1
22.7
25.6
28.8
32.5
37.2
43.5
53.0
69.1
80.7
94.6
108.9
126.4
114.8
84.7
60.6
258◦.5
263.5
268.2
272.8
277.2
281.5
285.4
286.9
277.1
256.5
219.8
189.8
201.3
184.7
104.0
70.5
15.7
15.2
14.6
14.0
13.3
12.4
11.4
10.2
8.7
7.9
7.0
6.1
5.0
6.1
7.6
9.1
19.7
19.0
18.2
17.3
16.3
15.1
13.6
11.9
9.7
8.5
7.0
5.3
3.4
5.5
8.2
10.6
S
′′
S
.31 +2
−2
−0
.8 +117
.5
−2.8 +116
−0.31 +2.7
−2.9 +115
−0.32 +2.8
−3.0 +113
−0.33 +3.0
−3.1 +110
−0.34 +3.2
−3.2 +107
−0.35 +3.4
−3.3 +94
−0.38 +3.5
−3.5 +71
−0.42 +3.2
−3.3 +30
−0.46 +1.5
−2.1
−0.36 +0.1
+8
−0.1
−0.03 +0.1
+1
−0.3
−4
+0.04 +1.7
+0.99 −8.7
+9.8 +204
+2.02 −7.4 +22.1 +219
+0.66 +17.4 +36.6 +508
−1.47 +4.7
−2.8 +723
′′
S
′′
.25 +10
−0
.3
−0.26 +10.3
−0.27 +10.4
−0.29 +10.4
−0.31 +10.4
−0.35 +10.3
−0.40 +10.0
−0.53
+8.9
−0.90
+5.6
−1.22
+1.7
−0.97
−3.4
−0.39
−5.9
+1.10 +19.9
+2.12 +22.3
+1.78 +44.0
−0.87 +55.6
in-depth orbital analysis of C/1945 X1 that is described
in the previous sections. With the observed arc of merely
four days and the images taken with small cameras of
poor spatial resolution and long exposures tracked at
sidereal rates, the data may at first sight seem overex-
ploited. However, the addressed objectives are not only
to learn as much as possible about the orbital behavior
of C/1945 X1 in the context of other Kreutz sungrazers,
but, as illustrated in Table 11, also to provide a basis for
constraining an ephemeris for the comet's motion both
back and forward in time from the observed period, a ca-
pability that is amply put to use in our search for more
images of the comet, as described in Section 7.
7. SEARCH FOR MORE DATA ON COMET C/1945 X1
The preceding sections demonstrate that, in spite of
some evidence to the contrary, the current level of our
knowledge of comet C/1945 X1 is less than sufficient for
addressing the issue of whether or not this object was
indeed a dwarf Kreutz sungrazer.
It is essential that,
where at all possible, existing information be rechecked,
reanalyzed, and reinterpreted, and that a vigorous search
for, and examination of, additional data be undertaken.
All our efforts will be aided substantially by products of
the Harvard College Observatory's DASCH Project, cur-
rently in progress, that we referred to in Section 3. As
for the new data, our proposed work has three objectives:
(i) to identify, astrometrically measure, and reduce any
further images of the comet in the collection of digitized
Boyden patrol plates (once they become available) with
the aim to refine the orbit by extending the arc covered
by observations as far back in time as possible; (ii) to
constrain the comet's physical behavior by deriving mag-
nitudes from all relevant digitized photographs and com-
pare the comet's light curve with that of other sungraz-
ers; and (iii) to search for and examine the comet's post-
perihelion images taken at Boyden for any traces of the
object or its debris, including the presence, appearance,
and orientation of a headless tail, diagnostic of the ob-
ject's fate near perihelion.
7.1. Search for Additional Images
Our extensive orbit examination in Section 6 allows
us now to extrapolate, with a fair degree of confidence,
the comet's motion way beyond the period of December
11 -- 15 and to realistically estimate uncertainties of an
ephemeris as a function of time. The nominal geocentric
ephemeris of the comet for the period of time from 1945
September 12.0 TT to 1946 January 30.0 TT, presented
in Table 11, is computed from the nongravitational solu-
tion N (N)
mod[0.05]{ℑ0} (whose elements and residuals are in
Table 10), but the differential ephemerides, derived from
the solution N (N)
std{ℑ0} (elements and residuals in Ta-
ble 9), from the gravitational solution G{ℑ0} (Table 7),
and from the orbit D′ (see the text near Table 6), are
also included for comparison. The two numbers listed in
Table 11 for each of these three additional solutions are
the corrections in right ascension and declination, respec-
tively, that are to be added to the nominal coordinates
in columns 2 and 3; note the different units used.
The first six columns of Table 11 are self-explanatory.
The seventh column gives the position angle of the pro-
longed radius vector (direction away from the Sun), and
the eighth and ninth columns include a predicted ap-
parent magnitude on two assumptions for the brightness
variations: as an inverse fourth and inverse seventh power
of heliocentric distance, with no phase effect. Both ver-
sions were normalized to the comet's reported magnitude
7 at discovery on December 11, which predicts a magni-
tude of 9.5 at 1 AU from both the Sun and the Earth
for the inverse fourth-power law and 10.6 for the inverse
seventh-power law (see Section 7.2).
We find the agreement among the ephemerides based
on the orbit D′ and the two nongravitational solutions
quite remarkable. On the other hand, the ephemeris from
the gravitational solution based on the same observations
as the two nongravitational orbits is very disappointing,
differing from the other three by as much as 0◦.2 over the
period of 4.5 months. We note that the three consistent
ephemerides come from the orbital sets that imply that
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
15
Additional Boyden Plates with Possible Images of Comet C/1945 X1.
Table 12
Plate
numbera
UT time at
mid-exposure
1945 Sept. 19.11997
RB 14049
Oct. 10.07683
AM 25110
17.07399
RB 14086
27.06607
AM 25132
Nov. 5.05881
AM 25145
12.06740
AM 25150
13.03559
RB 14121
15.06371
AM 25169
Dec. 7.04693
AM 25187
8.01580
AM 25190
AM 25208
14.03336
AM 25231e 1946 Jan. 5.04600
RB 14183e
6.04985
RB 14184f
8.07729
21.79113
AM 25240
21.84446
AM 25241
AM 25244
24.85635
30.85589
AM 25260
Plate centerb
R.A.
Dec.
Exp.
time
(min)
h
m
04
.6 −15◦17′
8
7 04.5 −15 09
8 04.0 −30 17
9 03.5 −45 24
9 04.2 −30 24
10 04.0 −45 29
8 03.2 −45 17
10 33.7 −60 31
11 04.6 −45 32
10 33.7 −60 31
16 40.9 −70 12
12 05.2 −70 33
12 05.2 −60 33
17 17.5 −47 19
0 05.1 −44 27
1 33.6 −59 29
5 25.4 −68 55
5 03.9 −29 52
60
85
120
60
90
80
120
85
90
90
40
90
115
40
90
60
90
90
Logbook
reference
rb10 148
am45b 0144
rb10 132
am45b 0148
am45b 0150
am45b 0152
rb10 138
am45b 0154
am45b 0156
am45b 0156
am45b 0158
am45b 0162
rb10 148
rb10 148
am45b 0164
am45b 0164
am45b 0164
am45b 0166
Comet's position (absolute and relative)b,c
R.A.
Dec.
Plate Dist.
P.A.
Edge Observerd
h
m
46
.9 −11◦13′ ON
5◦.91
7
8 11.4 −16 18
ON 16.11
8 20.3 −18 39
ON 12.20
8 34.4 −22 55 OFF 23.25
8 49.7 −28 04
ON
3.93
9 05.1 −33 25
ON 16.54
9 07.6 −34 16 OFF 16.51
9 13.4 −36 11 OFF 27.50
13 04.0 −65 58 OFF 25.94
13 34.2 −66 28
ON 20.50
16 34.1 −60 38
ON
9.60
18 32.4 − 42 24 OFF 52.65
18 38.1 −44 57 OFF 55.57
18 53.5 −50 19 OFF 16.02
1 36.3 −66 22 OFF 25.13
1 37.8 −66 15
6.78
2 40.0 −59 43 OFF 19.62
ON 23.86
3 38.4
-47 05
ON
312◦.6
96.3
18.5
342.8
305.4
311.8
53.8
323.1
152.4
126.1
355.0
112.7
121.9
109.7
158.5
176.5
277.3
217.8
6◦.85
1.18
2.33
0.71
13.99
4.87
2.12
0.50
1.49
0.63
11.94
31.37
35.99
3.88
1.89
14.73
2.26
2.56
Bl(?)
du Toit
J(?)
Bl(?)
Bester
J(?)
Bl(?)
J(?)
du Toit
J(?)
J(?)
Bl(?)
Bl(?)
Bester
Britz
Britz
Bl(?)
Bl(?)
a AM plates taken with the Cooke lens (covering 34◦.48 in R.A. and 43◦.10 in Dec.); RB plates with the Ross-Fecker camera (22◦.32 by 27◦.90).
b Equinox J2000.
c Column Plate indicates whether or not the comet's predicted position is within the plate limits; columns Dist. and P.A. show the predicted
position relative to the plate center; and in column Edge is the distance of the comet's image from the plate edge (either on or off the plate).
d The identities of the observers whose abbreviations are Bl and J could not be determined, major efforts notwithstanding; see also Table 1.
e This plate may show a segment of an early post-perihelion tail, if the comet still produced significant amounts of dust (see Figure 6 or 7).
f This is the only deliberate attempt at detecting C/1945 X1 after perihelion; unfortunately, the searched position is incorrect (see Figure 5).
the line of apsides lies within 0◦.01 of the reference value,
while the G{ℑ0} set leaves an offset of nearly 0◦.08.
To further appraise the magnitude of ephemeris uncer-
tainties, we completed 200 Monte Carlo runs with pure
gravitational solutions fitted to the {ℑ0} set of observa-
tions, assuming, conservatively, measuring errors of up
to ±9′′ in the December 11 -- 12 positions and up to ±2′′.5
in the December 14 -- 15 positions. The standard devia-
tions from the 200 clones vary, in the 1945 September --
November period, between 2′.4 and 3′ in right ascension
and between 1′.4 and 5′ in declination; in the 1946 Jan-
uary period, the range is from 6′ to nearly 20′ in right
ascension and from 8′ to 31′ in declination. On the av-
erage the offsets of the G{ℑ0}-based ephemeris in Table
11 amount to about 40% of these standard deviations.
To sum it up, we believe that our nominal ephemeris is
accurate, in a worst case scenario, to a few arcmin over
the entire period of interest to us.
The next step is a search for all patrol plates whose
fields cover (or could cover) the comet's ephemeris posi-
tion. Ideally, the physical size of each plate and its scale
provide the angular distance from the center to an edge
as a function of a position angle. It is in principle very
easy to establish whether the comet's image is or is not
located within the plate's limits. In reality, the problem
is less straightforward because of a potentially significant
error in the reported position of the plate center. Because
of this uncertainty (much larger than in the ephemeris),
the comet's predicted position may not be within the lim-
its of a plate when it should, or vice versa. The search
is further complicated by the comet's unknown bright-
ness, the observing conditions (the sky transparency, the
moon interference, the limiting magnitude, etc.), the ex-
posure time, and the distribution of field stars (needed
for the astrometry), all of which determines whether the
comet's image, even if present on the plate, can in fact be
recognized, astrometrically measured and reduced, and
photometrically evaluated.
Table 12 presents information on a total of 18 plates
that may show an image of C/1945 X1 or its debris,
including the unreduced plate AM 25208 from December
14 (Table 1) but not the five from Table 4. Because of the
position uncertainties, we include plates that nominally
miss the comet's position by up to 2◦.5.
These are all patrol plates, with very large fields of view
(Section 3). The Ross-Fecker plates (RB) cover a rectan-
gle of 22◦.32 in right ascension and 27◦.90 in declination
(∼600 square degrees), the Cooke plates (AM) 34◦.48 by
43◦.10 (∼1500 square degrees). The column of Table 12,
titled Plate, shows whether the comet's predicted posi-
tion, given by the standard polar coordinates relative to
the plate center, is (ON) or is not (OFF) within the limits
of the plate. The column Edge indicates the distance of
the comet's predicted position from the plate edge. The
smaller the number in an ON case, the more difficult it
will be to identify and astrometrically evaluate the im-
age because of optical imperfections far from the optical
axis of the instrument and a potential lack of appropriate
field stars. The smaller the number in an OFF case, the
greater is a chance that the image could, after all, appear
on the plate because of the positional uncertainties.
We list a total of seven preperihelion plates on which
the comet should show up if bright enough and four such
plates with the comet's predicted location barely missed;
see Section 7.2 for the brightness constraints. After peri-
helion there are only two potentially useful plates by the
16
Sekanina & Kracht
end of January to check whether any material part of the
head survived; two additional plates just miss the comet's
expected position. On one of the plates taken on January
21, we predict that the comet's location is not far from
the center. The comet's head should also appear on five
plates exposed during February (AM 25269 on the 7th,
AM 25272 and AM 25273 on the 8th, RB 14224 on the
19th, and RB 14246 on the 27th), but there is no point
in examining these unless the comet shows up, rather
unexpectedly, on the January plates.
Finally, there is a group of three plates (footnotes e and
f in Table 12), which miss the predicted position of the
comet's head by a wide margin but may be relevant to a
search for traces of the comet's dust debris (Section 8.2).
This group includes the January 8 plate (Table 1), taken
with the Ross-Fecker camera specifically for the purpose
to recover C/1945 X1 (Section 3). Another plate, taken
almost simultaneously with the Metcalf refractor (Ta-
ble 1), is not listed, because it is deemed useless on ac-
count of its small field.
7.2. Photometry and the Light Curve
To predict the brightness of a comet is always risky, as
has amply been documented by historical examples. For
the Kreutz sungrazers, the task is further complicated by
apparent differences between the shapes of the preperihe-
lion light curves of a prominent Kreutz comet (observed
as a bright object from the ground) and a dwarf sun-
grazer (observed only from SOHO/STEREO ), as sug-
gested by Knight et al. (2010). However, a preperihe-
lion light curve is known for only one prominent sun-
grazer, C/1965 S1 (Sekanina 2002).8 It is not advisable
to generalize this light curve as a standard attribute of
all bright Kreutz sungrazers. This caveat is supported by
the perceived differences among the post-perihelion light
curves for five prominent Kreutz objects (C/1843 D1,
C/1882 R1, C/1963 R1, C/1965 S1, and C/1970 K1),
which were fading with heliocentric distance at average
rates from r−3.3 to r−5.1 (Sekanina 2002).
Knight et al. (2010) concluded that the rate of bright-
ening of the dwarf Kreutz sungrazers observed with the
SOHO coronagraphs changes strikingly near 24 R⊙ (0.11
AU) from the Sun from ∼ r−7.3 to ∼ r−3.8; this slower
rate of brightening then holds, according to them, down
to 16 R⊙ (0.075 AU), where the light curve begins to
flatten. They estimated that the steep rate is unlikely to
persist at heliocentric distances greater than 50 R⊙ (0.23
AU) and that farther from the Sun the dwarf sungrazers
brighten on their way to perihelion by again following
an r−3.8 law. However, Ye et al. (2014) suggested that
there is a significant diversity in the preperihelion light
curves of the dwarf Kreutz sungrazers. They called at-
tention to either an outburst (whose amplitude exceeded
5 magnitudes) or an extremely steep brightening (r−11
or steeper) of a dwarf sungrazer C/2012 E2 between 1.06
AU and 0.52 AU from the Sun, while the object followed,
on the average, an ∼ r−4 (or flatter) law during its ap-
proach to perihelion between at least 0.52 AU (rather
8 Only a single, crude preperihelion magnitude estimate is avail-
able for C/1882 R1, the brightest Kreutz sungrazer over the past
two centuries. This estimate is based on reports of its rivaling
Venus in brightness at the time of discovery, 12 days before peri-
helion (Gould 1883).
than 0.23 AU) and 0.07 AU. Ye et al. also pointed out
that another dwarf Kreutz sungrazer, C/2012 U3, was
not detected at a heliocentric distance of 1.31 AU before
perihelion and was then much fainter than it should have
been if it followed the Knight et al.'s (2010) prediction.
Very illuminating is Ye et al.'s (2014) comparison of the
light curves of C/2012 U3 and C/2011 W3 (Lovejoy). Al-
though C/2011 W3 was not a dwarf sungrazer, it bright-
ened according to an r−6.9 law (Sekanina & Chodas 2012)
from 0.75 AU down to 0.34 AU, thus behaving very dif-
ferently from the prominent sungrazer C/1965 S1 (r−4.1
between 1.02 AU and 0.03 AU; Sekanina 2002). It ap-
pears that there is (i) a distinct possibility that, among
the dwarf sungrazers, the steep, r−7 law applies to helio-
centric distances of around 1 AU and possibly even fur-
ther away from the Sun; and (ii) at least some of these
objects appear to be subjected to outburst-like brighten-
ing at moderate distances from the Sun on their way to
perihelion.
The potential implications for the brightness evolution
of C/1945 X1 are obvious. A major obstacle to accept-
ing this object as a dwarf sungrazer is its considerable
brightness reported at the time of discovery. Compari-
son with C/2011 W3 shows that C/1945 X1 was intrin-
sically (i.e., after a correction has been applied for the
geocentric distance) fully 4.5 magnitudes (!) brighter at
the same heliocentric distance. Assuming the reported
estimate is correct, this problem can only be overcome
if one accepts that C/1945 X1 experienced a major out-
burst some time before the discovery, which was thereby
substantially facilitated. If the amplitude of the outburst
was greater than ∼4.5 magnitudes, and therefore compa-
rable to that of C/2012 E2, C/1945 X1 would have been
fainter than C/2011 W3 before the onset of the outburst.
From the standpoint of chances that C/1945 X1 could
be detected on any of the pre-discovery plates in Table
12, the pivotal parameter is the time when the putative
event took place. If this general scenario is correct, the
comet's unusual brightness is indeed explained, but there
is rather a slim chance of detecting the comet on any of
the pre-outburst plates.
An alternative explanation of the reported magnitude
of C/1945 X1, if it should be a dwarf sungrazer, is simply
a gross overestimation of its brightness by the discoverer.
In any case, these considerations show the critical need
for a thorough photometric examination of the comet's
Boyden images, both those from the period of December
11 -- 15 and any pre-discovery ones.
Keeping all our options open, we illustrate the dra-
matic differences, over a wide range of heliocentric dis-
tances, between the C/1945 X1 brightness predictions
based on the r−4 and r−7 laws by listing the apparent
magnitudes in columns 8 -- 9 of Table 11. One should,
however, be able to discriminate between the two laws
over a much shorter range of r.
In fact, in the known
Boyden images taken over the period of December 11 --
15, the comet should have brightened by 0.7 magnitude if
the r−4 law applied, but by 1.3 magnitudes if the steeper
law was in effect. The ephemeris suggests that C/1945
X1 might have been as bright as magnitude 15 in the sec-
ond half of September, about 3 months before perihelion.
In order that no image be missed, we began our search
for plates that fit the predicted position of the comet as
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
17
early as mid-September, since the Ross-Fecker camera,
the more powerful of the two patrol instruments, has
nominally a limiting magnitude 15 as well (Section 3).
8. WHAT KIND OF A SUNGRAZER WAS C/1945 X1?
Evidence of the fate of C/1945 X1 after its passing
through perihelion is vital for determining the place of
this object in the hierarchy and evolution of the Kreutz
system and for answering the question of whether it was
a dwarf sungrazer. We propose to establish this from
comparison with some better investigated sungrazers.
8.1. C/1945 X1 and the Transition Sungrazers
Based on large numbers of observations, the members
of the Kreutz system are usually divided into two cate-
gories: prominent (bright and often quite spectacular as
seen from the ground) and dwarf (defined in Section 1).
This is a very simplistic classification, because it ignores
intermediate objects, which occupy the transition be-
tween both categories and, albeit scarce, are momentous
for a better understanding of the disintegration process
of the Kreutz sungrazers.
In an effort to determine where C/1945 X1 is likely to
fit in, we selected three "standards" to define a scale for
transition objects in the order of increasing similarity to
the dwarf members. The standards were chosen to be
represented by C/2011 W3, C/1887 B1, and C/2007 L3.
Next we describe their diagnostic properties, which will
be searched for on the best suited plates of C/1945 X1
in order to classify the comet on this scale.
C/2011 W3 was the most persistent and nearest the
status of a prominent member among the three objects.
With a perihelion distance of 1.19 R⊙, this comet sur-
vived intact (without its nucleus disintegrating com-
pletely) until 38 ± 5 hours after perihelion, at which time
it suddenly fell apart in a terminal outburst , losing sud-
denly its residual, ∼1012 g nucleus and head (Sekanina
& Chodas 2012). The tail, made up of microscopic dust
particles ejected from the nucleus before perihelion, was
seen to survive perihelion for at least 24 hours as a syn-
dyname -- a locus of dust released at different times but
subjected to the same radiation pressure acceleration β
-- of β = 0.6 the solar gravitational acceleration, typi-
cal for dielectric submicron-sized particles. Only grains
ejected less than 0.1 day before perihelion, which ap-
proached the Sun within 1.8 R⊙, sublimated away. After
perihelion the comet formed a new, spectacular, head-
less tail, most of which was a product of the event ∼38
hours after perihelion. This tail was a synchrone -- a lo-
cus of dust grains of different sizes (subjected to a broad
range of radiation pressure accelerations), all released at
the same time -- that was under observation for up to
90 days, until mid-March 2012. Its representative length
during most reported sightings again suggested a maxi-
mum radiation pressure acceleration of β ≃ 0.6 the solar
gravitational acceleration. Only during the late obser-
vations (60 -- 90 days after perihelion) did the visible tail
consist of merely micron-sized and larger grains, whose
β ≪ 0.6 the solar gravitational acceleration.
C/1887 B1 is the runner-up to C/2011 W3. Avail-
able information is rather limited, because this object
was discovered only after perihelion as a headless, nar-
row, ribbon-like tail, again a synchrone. Nonetheless, the
systematic variations in this tail's orientation with time
suggest that the nucleus survived intact until 5.8 ± 0.8
hours after perihelion (Sekanina 1984), at which point it
must have suddenly disintegrated in a terminal outburst
similar to that of C/2011 W3. At perihelion the comet
was merely 1.04 R⊙ from the Sun (Sekanina 1978; Seka-
nina & Chodas 2004). The tail, truly spectacular during
the early sightings from 8.5 days after perihelion on, was
under observation until 18.5 days after perihelion and its
length implied a maximum radiation pressure accelera-
tion of dust particles that dropped from ∼0.4 the solar
gravitational acceleration at discovery to ∼0.05 during
the last three reported observations. The minimum par-
ticle size in the visible part of the tail was thus increasing
with time from less than 1 micron to several microns.
C/2007 L3 is, of the three standards, the most re-
mote from the prominent Kreutz-system members, be-
ing in fact a representative of a relative small group of
bright dwarf Kreutz sungrazers, which develop long, ex-
tremely narrow tails upon their approach to perihelion
and which we refer to hereafter as the superdwarfs. Like
ordinary dwarf sungrazers, their nuclei fail to survive per-
ihelion, so that they form no post-perihelion tails. We
chose C/2007 L3 as a representative primarily because it
was imaged by both SOHO and STEREO , unlike most
of its peers. The comet had a perihelion distance of
1.530 R⊙ (Marsden 2008), very close to our best esti-
mate for C/1945 X1, its brightness peaked at magnitude
∼3 and its head disappeared about 80 minutes before
perihelion, at a heliocentric distance of 2.75 R⊙ (Green
2007). The SOHO and STEREO images were studied
in considerable detail by Thompson (2009), who closely
confirmed the earlier conclusions by Sekanina (2000),
based on several objects, that preperihelion images of
these tails show that they are synchrones, referring to
emission-shutoff times of 26 ± 7 hours before perihelion
and to heliocentric distances at shutoff of 23.5 ± 4.5 R⊙.9
This average agrees remarkably well with Knight et al.'s
(2010) distance at which the slope of the light curves sud-
denly changes from −7.3 to −3.8 (Section 7.2).10 The
tail of C/2007 L3 survived the comet's perihelion, as did
the preperihelion tail of C/2011 W3. Thompson exam-
ined its remnant and found, unlike Sekanina & Chodas
(2012) for C/2011 W3, that the feature remained syn-
chronic to its last examined image, nearly 17 hours af-
ter perihelion, but that its position was slightly mod-
ified by a loss of angular momentum, possibly due to
atmospheric drag from the solar corona. The appar-
ent discrepancy is explained by the fact that the dust
emission of C/2011 W3 did not shut off about 1 day
before perihelion, but continued. Both Sekanina (2000)
and Thompson (2009) detected the maximum radiation
pressure accelerations of β ≃ 0.6 the solar gravitational
attraction. Comet C/2007 L3 was not the only Kreutz
superdwarf whose dust tail was observed to survive per-
ihelion. Other examples are C/1979 Q1, the first Sol-
wind comet (Michels et al. 1982), C/1998 K10, one of
the dozen comets examined by Sekanina (2000), etc. It is
9 For the best two imaged comets (C/1996 Y1 and C/1998 K10)
in his set, Sekanina (2000) noticed that the tail was actually bent
and, unlike its bright synchronic portion, the faint, outer part was
a syndyname with a range of the ejection times.
10 This evidence may be interpreted to indicate that dust emis-
sion essentially ceases at 24 R⊙ from the Sun and the r−3.8 law of
brightening measures an effect of the object's fragmentation.
18
Sekanina & Kracht
likely that most, if not all, superdwarfs would show their
tails surviving perihelion, if their images are closely ana-
lyzed. About 10 such comets from 1995 -- 2005 in Knight
et al.'s (2010) Table 2 and all 19 sungrazers from 2006 --
2013 in Sekanina & Kracht's (2013) Table 1 belong to
this subcategory, which is at present estimated to include
about three dozen objects.
So where does C/1945 X1 fit in? Employing the con-
straint that no tail was ever reported with the unaided
eye after perihelion, it is certain that this comet was too
faint to be placed between C/2011 W3 and C/1887 B1
or even near C/1887 B1. Since a preperihelion tail is
always found to survive for only 1 day or so after peri-
helion, there is no chance of detecting it on any of the
plates listed in Table 12. Should we find traces of the
comet's synchronic tail, a product of a post-perihelion
outburst, on any of these plates, then C/1945 X1 is to
be classified on this three-point scale between C/1887 B1
and C/2007 L3, in which case it was a transition object
more massive than a superdwarf.
If not, it has to be
concluded that C/1945 X1 cannot be positioned higher
than C/2007 L3, so that, at best, it was a superdwarf.
8.2. Search for Traces of C/1945 X1 After Perihelion
From these considerations it follows that while a search
for, and examination of, the comet's preperihelion im-
ages on the plates listed in Table 12 should serve to im-
prove our understanding of the comet's orbital motion
and its light curve, a careful search for traces of the comet
on, and examination of, the plates taken after perihelion
should provide information on the comet's fate and help
answer the question of whether C/1945 X1 was a transi-
tion object or a dwarf sungrazer; because of the lack of
information on the comet's appearance in close proximity
of perihelion (Table 12), we cannot distinguish between
its classification as a dwarf or a superdwarf.
Before we discuss specific search opportunities, we note
that the dust released from this comet before perihelion
was affected by sublimation less severely than the dust
from C/2011 W3. The orbital elements for C/1945 X1's
preperihelion ejecta subjected to radiation pressure ac-
celerations β = 0.6 the solar gravitational acceleration
are summarized in Table 13.
It follows that if this
comet's dust had sublimation properties similar to those
of the dust from C/2011 W3, only particles ejected within
∼1 hour of perihelion should have sublimated completely,
because of a greater perihelion distance of C/1945 X1.
A major delay of several days in passing through perihe-
lion is noted as a yet another effect: for example, parti-
cles ejected 50 days before the comet's perihelion did not
reach their own perihelion until late on January 5.
As to the sources of information on comet C/1945 X1
after perihelion, a remote opportunity occurred on 1946
January 3.51 UT, the time of a solar eclipse.11 Unfor-
tunately, it was merely partial, of magnitude 0.553, and
visible only from Antarctica and surrounding oceanic re-
gions. Even less of the Sun was occulted at the locations
of several permanent British scientific stations in Antarc-
tica near the southernmost tip of South America (Oper-
ation Tabarin; e.g., Haddelsey 2014). At J. Shanklin's
suggestion we checked the UK Met Office Archives in
11 See the website http://eclipse.gsfc.nasa.gov/SEplot1901/
SE1946Jan03P.GIF.
Table 13
Orbital Elements of Dust Particles Subjected to β = 0.6
and Ejected from Comet 1945 X1 up to Perihelion.
Time of
ejection
(days)a
Time of
perihelion
(days)a
−50
−40
−30
−25
−20
−15
−10
−7
−5
−3
−2
−1
−0.7
−0.5
−0.3
−0.2
−0.1
−0.05
−1h
0.0
+8.868
+7.142
+5.410
+4.540
+3.667
+2.788
+1.899
+1.357
+0.989
+0.610
+0.412
+0.203
+0.136
+0.089
+0.041
+0.017
−0.004
−0.010
−0.009
0.000
Perihelion
distance
Orbital
eccentricity
(R⊙)
3.757
3.749
3.732
3.721
3.706
3.682
3.641
3.596
3.545
3.450
3.356
3.134
3.027
2.896
2.673
2.481
2.144
1.843
1.779
1.530
1.035
1.041
1.049
1.055
1.064
1.077
1.101
1.127
1.157
1.217
1.279
1.426
1.527
1.642
1.861
2.082
2.567
3.150
3.299
4.000
a Reckoned from the time of the comet's perihelion passage.
Exeter and, not surprisingly, found no report on any rel-
evant observations at the time.
Back at Boyden, an attempt was made by M. J. Bester
on January 8 to recover the comet; the search, as pointed
out (Sections 3 and 7.1), was conducted with the Metcalf
refractor and the Ross-Fecker piggyback camera. Both
exposures pointed in the same direction because they
overlapped each other; the Metcalf 45 minute exposure
began 20 minutes earlier and was longer by 5 minutes.
The pointing of the telescope is, however, a mystery: as
performed, it could not accommodate the comet's pre-
dicted position within the limits of either plate.
January 8 was the date of publication of Cunningham's
(1946b) ephemeris, but early in the morning Bester could
not have been in the possession of the official announce-
ment from Copenhagen. Yet, the telescope was set to
point precisely in accord with the ephemeris in decli-
nation and was -- most peculiarly -- off by exactly
2 hours to the west of the ephemeris place in right as-
cension. Such a coincidence surely cannot be accidental.
We presume that Bester was notified directly by Cun-
ningham ahead of the ephemeris' publication, perhaps
by cable, but there was a miscommunication, an error
in one coordinate.
In any event, the time of Bester's
observation was fundamentally incorrect because during
the exposures the comet, less than 30◦ from the Sun,
was below and, later, very close to the horizon. Since
the comet was almost exactly south of the Sun, it could
have been observed either shortly after sunset or, some-
what preferably, just before sunrise. The need to take
a fairly long exposure did of course constrain the choice
of the observation time. If observed at the correct time
with the telescope set to point in accord with Cunning-
ham's ephemeris in both coordinates, the comet's po-
sition would have been exposed on both plates, even
though only ∼0◦.5 from the edge of the Metcalf plate.
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
19
PUTATIVE DUST TAIL OF C/1945 X1
ON BOYDEN EXPOSURE RB 14184
TAKEN ON 1946 JANUARY 8
✎☞♠♥SUN q
✍✌
··········································❞ -- 40
·························································································
-- 0.5
······················································································❞
-- 0.5 d
β = 0.6
q
q
q
q
q
q
q
q
q
q
q
q
COMET
MOTION
························································❞ -- 1.2
·························
-- 1.2 d
0.45
t
PLATE
CENTER
PLATE RB 14184
··········································❞ -- 2.5
·················································❞ -- 20
·················································❞ -- 10
···················································································❞ -- 5
···········································································t0.35
···········································································t0.5
······································································································································································································································································································································································································································································································································································
×
q✉
····································································································t0.2
···············································································································································································t0.07
········································································································································································································t◗◗
····································································································t✑✑ 0.04
························································································································t
····································································································t
······················································································································································t
···········································································t
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣
♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣♣❄❄
···························································································································································t0.14
····································································································t
···········································································t
❆❆
····································································································t
························································································································t
❆❆
········································································································································································································t0.4
◗◗
0.02
············································································································································t0.12
❇
❇
0.06
····························································
····································································································
··························································································t
···········································································t
····································································································t
············································································································································t0.21
0.06
······························································
+0.5 d
···············
+2 hr
t0.30
0.03
0.08
t0.15
-- 5 hr
·································································································
·······························································
-- 2 hr
0.12
0.09
··········································
-- 1 hr
+1.6 d
0.6
0.2
0.06
0.04
1.2
tπ
A
N
G
U
L
A
R
D
I
S
T
A
N
C
E
F
R
O
M
C
O
M
E
T
I
N
D
E
C
L
I
N
A
T
I
O
N
+30◦
+20◦
+10◦
0◦
-- 10◦
-- 20◦
-- 30◦
+10◦
0◦
-- 10◦
-- 20◦
-- 30◦
-- 40◦
-- 50◦
ANGULAR DISTANCE FROM COMET IN RIGHT ASCENSION
Figure 5. The predicted projected position of C/1945 X1 relative to the sky area exposed on the plate RB 14814, taken by M. J. Bester
with the Ross-Fecker camera on 1946 January 8.07729 UT. While the comet's head was more than 4◦ off the eastern edge of the plate,
much of the dust tail made up of particles released near and shortly before perihelion occupy the lower, southern part of the plate's field.
Dust ejected as far back as ∼40 days before perihelion occupies the rest of the plate. The curves to the south of the plate center are
the synchrones, the loci of the dust that was ejected from the comet at particular times between 1.2 days before perihelion and 1.6 days
after perihelion; the curve tπ is the perihelion synchrone. The numbers along the synchrones are the values β, the ratio between the solar
radiation pressure acceleration and the solar gravitational acceleration; β and the ejection time govern the motions of the dust ejecta.
Values near β ≈ 0.1 are typical for micron-sized grains, those near or in excess of 0.5 refer to a variety of submicron-sized grains. The thick
curve at the top, labeled β = 0.6 is a syndyname defined by this β ratio and typical for dielectric submicron-sized particles. Shown along
this curve are the times of ejection in days before perihelion; see text for the significance of this syndyname.
Figure 5 displays the projected positions of the comet
and the Sun relative to the pointing direction of the plate
RB 14184, taken with the Ross-Fecker camera on Jan-
uary 8, 11 days after perihelion. Although dust ejecta are
predicted to be all over the plate, their detection, contin-
gent on the time-dependent emission rate, is problematic
on account of their generally low surface brightness, ex-
cept when released during a sharply peaked event, such
as a terminal outburst. The curves emanating in the
figure from the comet's head and distributed mostly in
the lower part of the plate are the near-perihelion syn-
chrones. Their selected range is from 1.2 days before per-
ihelion to 1.6 days after perihelion. Few post-perihelion
ejecta project onto the plate, barely touching its south-
eastern corner, and only a limited fraction of the perihe-
lion ejecta crosses the plate. On the other hand, much
of the dust released as far back as ∼40 days before per-
ihelion is projected onto the plate. We emphasized in
Section 8.1 that the preperihelion tails of the transition
object C/2011 W3 and of C/2007 L3 and many other
(if not all) superdwarfs survived perihelion for not much
longer than ∼1 day at the most. Such features are too
close to the Sun to observe from the ground. It is only for
the sake of interest that Figure 5 shows the syndyname
β = 0.6, made up of dielectric submicron-sized particles
ejected from C/1945 X1 before perihelion, to miss the
field of the sky covered by the plate RB 14184, nearly
grazing it along the entire northern edge.
The plate RB 14184 may prove useful. If C/1945 X1
experienced a terminal outburst and disintegrated within
an hour or two after perihelion and if the ejecta contained
a large abundance of grains several microns in diameter,
there is a possibility that this debris could show up as
a band of light across the lower left corner of the plate.
The phase angles favor a modest forward-scattering ef-
fect, being mostly in a range of 100◦ -- 110◦. Submicron-
sized ejecta released from the comet at the same times
should miss the plate. Thus, a detection or nondetection
of a post-perihelion synchronic tail could provide mean-
ingful constraints on the existence of a terminal outburst
and on the time of nucleus disintegration.
We found two more plates that might provide addi-
tional limited information on a potential post-perihelion
terminal outburst: AM 25231, taken with the Cooke lense
20
Sekanina & Kracht
BOYDEN EXPOSURE AM 25231
TAKEN ON 1946 JANUARY 5
BOYDEN EXPOSURE RB 14183
TAKEN ON 1946 JANUARY 6
hr
6 hr
······················································
❡
1.05
3 1
2
·····························
❡
0.60
································································································
❡
0.90
2 hr
························
❡
············
0.34
·········································································································
❡
·················································································································
····························
❡
0.50
0.75
······································
×
PLATE
CENTER
A
N
G
U
L
A
R
D
I
S
T
A
N
C
E
I
N
D
E
C
L
I
N
A
T
I
O
N
+20◦
+15◦
+10◦
+5◦
0◦
-- 5◦
-- 10◦
-- 15◦
-- 20◦
A
N
G
U
L
A
R
D
I
S
T
A
N
C
E
I
N
D
E
C
L
I
N
A
T
I
O
N
+12◦
+9◦
+6◦
+3◦
0◦
-- 3◦
-- 6◦
-- 9◦
-- 12◦
······································································································································
❡
0.70
··········································
❡
······························································································
❡
0.50
0.30
0.45
0.33
2 hr
·········································
❡
···············································································
❡
··························································
··········································································································
❡
································································································································································
···························································································
··················································································································································
3 hr
0.60
0.40
❡
5 hr
·························································································································
❡
0.80
×
PLATE
CENTER
+15◦
+10◦
+5◦
0◦
-- 5◦
-- 10◦
-- 15◦
ANGULAR DISTANCE IN RIGHT ASCENSION
+9◦
+6◦
+3◦
0◦
-- 3◦
-- 6◦
-- 9◦
ANGULAR DISTANCE IN RIGHT ASCENSION
Figure 6. Predicted positions, on a Cooke plate AM 25231 taken
on January 5.04600 UT, of synchrones that depict C/1945 X1's
dust ejecta from hypothetical terminal outbursts 2, 3.5, and 6 hours
after perihelion. The smaller-font digits refer to particles subjected
to a variety of radiation pressure accelerations β between 0.34 and
1.05 the solar gravitational acceleration. Since the Kreutz comets
appear to have β ≤ 0.6, only outbursts that occurred between ∼2
and ∼4 hours after perihelion might be detected on this plate.
Figure 7. Predicted positions, on a Ross-Fecker plate RB 14183
taken on January 6.04985 UT, of synchrones depicting the comet's
dust ejecta from hypothetical terminal outbursts 2, 3, and 5 hours
after perihelion. The smaller-font digits refer to particles subjected
to a variety of radiation pressure accelerations β between 0.3 and
0.8 the solar gravitational acceleration. Since the Kreutz comets
appear to have β ≤ 0.6, only outbursts that occurred between ∼1
and ∼5 hours after perihelion might be detected on this plate.
in the morning of January 5, and RB 14183, taken with
the Ross-Fecker camera 24 hours later. They both are
listed in Table 12. The synchrones for the January 5
plate are depicted in Figure 6, those for the January 6
plate in Figure 7. The positions of dust particles rela-
tive to the comet's nucleus, determined by a computer
code in terms of their polar coordinates, were converted
to the polar coordinates of the offsets from the plate cen-
ter, using a technique described in some detail in the Ap-
pendix. An advantage of these plates relative to that on
RB 14184 is their exposure times, 115 and 90 minutes, re-
spectively, as opposed to only 40 minutes (Table 12). The
January 5 plate might provide useful information only if
the terminal outburst took place between, at most, ∼2
and ∼4 hours after perihelion; a synchronic tail contain-
ing earlier ejecta would miss this plate, while segments of
later synchrones limited to β > 0.6 the solar gravitational
acceleration are unlikely to refer to any real tails (Sec-
tion 8.1). Similarly, the January 6 plate might provide
information only on an outburst that occurred between
∼1 and ∼5 hours after perihelion. No forward-scattering
effect can be expected in either case, as the phase angles
range between about 60◦ and 80◦.
In the highly unlikely case that a major fragment of
the original nucleus survived for weeks after perihelion,
the best search opportunity is offered by a Cooke plate
AM 25241, taken on January 21, on which the comet is
predicted to be less than 7◦ from the center (Table 12).
Less favorable circumstances should accompany a search
on a plate AM 25260, taken 9 days later, on which the
comet's predicted position is only 2◦.5 from the edge. Ad-
ditional search opportunities are more discouraging still.
9. CONCLUSIONS
Comet C/1945 X1 has been rather an oddball among
the Kreutz sungrazers observed from the ground ever
since its discovery 70 years ago. Although the dilatory
handling of the Boyden plates contributed to the snail
pace in getting out the facts about the object, it was
nonetheless its apparent lack of luster near and after per-
ihelion that is primarily responsible for our ignorance of
its properties, fate, and place in the Kreutz system.
The comet's disappointing post-perihelion performance
-- especially the absence of a prominent dust tail -- pro-
vides a strong argument against its being on a par with
the headless sungrazers C/2011 W3 and C/1887 B1 that
survived perihelion with a fairly massive nucleus intact
but disintegrated shortly afterwards. The plausible sce-
narios for C/1945 X1 are thus limited to two: either the
comet's surviving mass was sufficient to generate a mod-
est, but not spectacular, post-perihelion tail, or the comet
completely disintegrated still before perihelion. While
there is not much of a difference between the two scenar-
ios, we use the second one to define a dwarf Kreutz sun-
grazer. The existence of a terminal outburst is therefore
critical to distinguish between the two options.
Taking account of the indirect planetary perturbations
on the premise that a precursor of C/1945 X1 separated
from its common parent with C/1882 R1 (Sekanina &
Chodas 2004), we determine that the line of apsides of
C/1945 X1 should point toward an ecliptical longitude
of Lπ = 282◦.44 and latitude of Bπ = +35◦.16 (Equinox
J2000) and that the comet's osculating semimajor axis a
should be 96.70 AU at an epoch 1945 December 28.0 TT
or 96.67 AU at an epoch December 11.0 TT.
Comet C/1945 X1 (du Toit) -- a Dwarf Kreutz Sungrazer?
21
Employing this value of a as an orbital constraint,
our Best Fit purely gravitational solution based on all
five observations left an unacceptably large offset of
0◦.57 from the nominal apsidal-line direction, apparently
because the third observation was inferior. This was
confirmed by four-observation gravitational solutions, of
which the one based on the positions from December 11,
12, 14, and 15 was by far the most promising, leaving
an offset from the nominal apsidal-line direction of 0◦.08,
much better but still not entirely satisfactory.
Applying a technique introduced recently by us in Pa-
per 1, we show that the incorporation of a nongravita-
tional acceleration of the orbital motion improves the fit
in terms of both the apsidal-line offset and mean resid-
ual. Use of Marsden et al.'s (1973) standard formalism
suggests an acceleration on the order of 10−6 AU day−2
(with a relative error of ±8 -- 17%), near the lower end
of a range typical for the dwarf sungrazers. A prefer-
able nongravitational model, based on a much steeper
acceleration law (apparently common among the dwarf
sungrazers; see Paper 1) than is the standard law, leads
to the same magnitude of the integrated effect but fits
the observations better, with the relative errors of the
nongravitational parameters reduced to ±5 -- 7%.
In order to augment the current, very limited data set,
an extensive search for further possible images of the
comet should be undertaken. We hope to be able to
conduct such a search, based on our orbital computa-
tions suggesting that at least 7 and perhaps as many
as 11 preperihelion Boyden patrol plates taken between
late September and mid-December may contain such im-
ages. The plates will be inspected in the near future, once
their digitized copies are made available by the DASCH
Project. At present we do not anticipate to expand our
search to archives of wide-field plates from other obser-
vatories (e.g., Tsvetkov & Tsvetkova 2012).
In addition to positional data of the comet, its pho-
tometry will also have to be performed to reexamine the
reported brightness at discovery and to settle the issues
of the rate of brightness variations with heliocentric dis-
tance and the chance of an early preperihelion outburst.
Comparison with C/2011 W3 and C/2012 E2, among
others, may help settle some of the issues of preperihe-
lion activity of the Kreutz sungrazers.
Finally, to pursue a solution to the complex problem
of C/1945 X1's place in the hierarchy of the Kreutz
system and the comet's fate, we also will examine some
post-perihelion Boyden patrol plates to search for both
potential relics of the comet itself and/or possible traces
of its dust tail as a product of a modest terminal out-
burst that may have occurred just hours past perihelion.
Whereas it appears that, masswise, C/1945 X1 could not
compete with either C/2011 W3 or C/1887 B1, we will
try to disentangle a mystery of whether it still could be
considered a transition object distinctly more massive
than a superdwarf, such as C/2007 L3, C/1998 K10,
or C/1979 Q1, or it does not differ materially from the
Kreutz dwarfs or superdwarfs -- and thus answer the
cardinal question posed in this paper's title.
The authors thank D. van Jaarsveldt, J. Grindlay, A.
Doane, and E. Los for their information on the observers
and the plate collection from the Boyden Station and
on the Harvard College Observatory's DASCH Project.
This research was carried out in part at the Jet Propul-
sion Laboratory, California Institute of Technology,
under contract with the National Aeronautics and Space
Administration.
APPENDIX
TRANSFORMATION OF DUST PARTICLE POSITION FROM
COMETOCENTRIC TO PLATE CENTER COORDINATES
The code computing the position of a dust particle, which
is a function of the ejection time, teject, and a ratio of
the radiation pressure acceleration to the solar gravita-
tional acceleration, β, provides the polar coordinates in
reference to the comet's nucleus in projection onto the
plane of the sky -- the angular distance from the nu-
cleus, D∗ (teject, β), and the position angle, Π∗ (teject, β),
reckoned from the north through the east. The equatorial
coordinates of the nucleus, right ascension α0 and decli-
nation δ0, and the coordinates of a particle, α∗ (teject, β)
and δ∗ (teject, β), are related to D∗ and Π∗ by the well-
known expressions,
cos D∗ = sin δ0 sin δ∗ + cos δ0 cos δ∗ cos ∆α∗ ,
cot Π∗ = cos δ0 tan δ∗ csc ∆α∗ − sin δ0 cot ∆α∗ , (19)
where ∆α∗ = α∗−α0. To convert the particle's position
from the cometocentric coordinate system to the coordi-
nate system centered on the center of a plate, we first
particle from the formulas
compute right ascension α∗ and declination δ∗ of the
sin2D∗ tan Π∗ (sin δ0 + cotD∗ cos δ0 sec Π∗ )
tan ∆α∗ =
cos2D∗ (1+ cos2δ0 tan2Π∗ )− sin2δ0
,
(20)
and
tan δ∗ = cos ∆α∗ tan δ0(1+cot Π∗ tan ∆α∗ csc δ0). (21)
Since the sign of cos ∆α∗ has a direct effect on the sign
of tan δ∗ and therefore also on the signs of sin δ∗ and
cos D∗, the quadrant of ∆α∗ in Equation (20) needs to
be chosen such that after inserting the values of ∆α∗ and
δ∗ from (21) into the first equation of (19) one gets the
correct value of D∗ and not its supplement.
Next, identifying α0 and δ0 with the coordinates of
the plate center, rather than the comet's nucleus, and
inserting them together with the particle's coordinates
α∗ and δ∗ into Equations (19), one obtains an angular
distance D∗ and a position angle Π∗ that determine the
particle's position relative to the plate center.
REFERENCES
Brueckner, G. E., Howard, R. A., Koomen, M. J., et al. 1995, Sol.
Phys., 162, 357
Cooper, T. P. 2003, MNASSA, 62, 170
Cooper, T. P. 2005, MNASSA, 64, 118
Cunningham, L. E. 1946a, HAC, 733
Cunningham, L. E. 1946b, IAUC, 1025
Gould, B. A. 1883, Astron. Nachr., 104, 129
Green, D. W. E. 2007, IAUC, 8883
Grindlay, J., Tang, S., Los, E., & Servillat, M. 2012, IAU Symp.,
285, 29
Haddelsey, S. 2014, Operation Tabarin: Britain's Secret Wartime
Expedition to Antarctica 1944 -- 1946. History Press, Stroud, UK
Hockey, T. 2009, BAAS, 41, 572
Howard, R. A., Moses, J. D., Vourlidas, A., et al. 2008, Space Sci.
Rev., 136, 67
22
Sekanina & Kracht
Knight, M. M., A'Hearn, M. F., Biesecker, D. A., et al. 2010, AJ,
139, 926
Kreutz, H. 1888, Publ. Sternw. Kiel, 3
Kreutz, H. 1891, Publ. Sternw. Kiel, 6
Kreutz, H. 1901, Astron. Abh., 1, 1
Marsden, B. G. 1967, AJ, 72, 1170
Marsden, B. G. 1989, AJ, 98, 2306
Marsden, B. G. 2008, MPEC 2008-G44
Marsden, B. G., & Williams, G. V. 2008, Catalogue of Cometary
Orbits 2008, p. 108 (17th ed.; Cambridge, MA: Smithsonian
Astrophysical Observatory, 195pp)
Marsden, B. G., Sekanina, Z., & Yeomans, D. K. 1973, AJ, 78, 211
Michels, D. J., Sheeley, N. R., Howard, R. A., & Koomen, M. J.
1982, Science, 215, 1097
Nagahara, H., Mysen, B. O., & Kushiro, I. 1994, Geochim.
Cosmochim. Acta, 58, 1951
Paraskevopoulos, J. S. 1945, IAUC, 1024
Pasachoff, J. M., Rus´ın, V., Druckmuller, M., et al. 2009, ApJ, 702,
1297
Sekanina, Z. 1978, QJRAS, 19, 52, 57
Sekanina, Z. 1982, AJ, 87, 1059
Sekanina, Z. 1984, Icarus, 58, 81
Sekanina, Z. 2000, ApJ, 545, L69
Sekanina, Z. 2002, ApJ, 566, 577
Sekanina, Z., & Chodas, P. W. 2002, ApJ, 581, 760
Sekanina, Z., & Chodas, P. W. 2004, ApJ, 607, 620
Sekanina, Z., & Chodas, P. W. 2007, ApJ, 663, 657
Sekanina, Z., & Chodas, P. W. 2012, ApJ, 757, 127 (33pp)
Sekanina, Z., & Kracht, R. 2013, ApJ, 778, 24 (13pp)
Sekanina, Z., & Kracht, R. 2015, ApJ, 801, 135 (19pp)
Simcoe, R. J., Grindlay, J. E., Los, E. J., et al. 2006, in Applica-
tions of Digital Image Processing XXIX, ed. A. G. Tescher, Proc.
SPIE, vol. 6312, 631217 (Bellington, WA).
Thompson, W. T. 2009, Icarus, 200, 351
Tsvetkov, M., & Tsvetkova, K. 2012, IAU Symp., 285, 417
van Heerden, H. J. 2008, MNASSA, 67, 116
Ye, Q.-Z., Hui, M.-T., Kracht, R., & Wiegert, P. A. 2014, ApJ,
796, 83 (8pp)
|
1302.3655 | 1 | 1302 | 2013-02-15T01:10:00 | Evidence for a Snow Line Beyond the Transitional Radius in the TW Hya Protoplanetary Disk | [
"astro-ph.EP"
] | We present an observational reconstruction of the radial water vapor content near the surface of the TW Hya transitional protoplanetary disk, and report the first localization of the snow line during this phase of disk evolution. The observations are comprised of Spitzer-IRS, Herschel-PACS, and Herschel-HIFI archival spectra. The abundance structure is retrieved by fitting a two-dimensional disk model to the available star+disk photometry and all observed H2O lines, using a simple step-function parameterization of the water vapor content near the disk surface. We find that water vapor is abundant (~10^{-4} per H2) in a narrow ring, located at the disk transition radius some 4AU from the central star, but drops rapidly by several orders of magnitude beyond 4.2 AU over a scale length of no more than 0.5AU. The inner disk (0.5-4AU) is also dry, with an upper limit on the vertically averaged water abundance of 10^{-6} per H2. The water vapor peak occurs at a radius significantly more distant than that expected for a passive continuous disk around a 0.6 Msun star, representing a volatile distribution in the TW Hya disk that bears strong similarities to that of the solar system. This is observational evidence for a snow line that moves outward with time in passive disks, with a dry inner disk that results either from gas giant formation or gas dissipation and a significant ice reservoir at large radii. The amount of water present near the snow line is sufficient to potentially catalyze the (further) formation of planetesimals and planets at distances beyond a few AU. | astro-ph.EP | astro-ph | Draft version January 13, 2021
Preprint typeset using LATEX style emulateapj v. 5/2/11
3
1
0
2
b
e
F
5
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
5
6
3
.
2
0
3
1
:
v
i
X
r
a
EVIDENCE FOR A SNOW LINE BEYOND THE TRANSITIONAL RADIUS IN THE TW HYA
PROTOPLANETARY DISK
K. Zhang1, K. M. Pontoppidan 2, C. Salyk3, and G. A. Blake4
Draft version January 13, 2021
ABSTRACT
We present an observational reconstruction of the radial water vapor content near the surface of
the TW Hya transitional protoplanetary disk, and report the first localization of the snow line during
this phase of disk evolution. The observations are comprised of Spitzer -IRS, Herschel -PACS, and
Herschel -HIFI archival spectra. The abundance structure is retrieved by fitting a two-dimensional
disk model to the available star+disk photometry and all observed H2O lines, using a simple step-
function parameterization of the water vapor content near the disk surface. We find that water vapor
is abundant (∼ 10−4 per H2) in a narrow ring, located at the disk transition radius some 4 AU from
the central star, but drops rapidly by several orders of magnitude beyond 4.2 AU over a scale length
of no more than 0.5 AU. The inner disk (0.5-4 AU) is also dry, with an upper limit on the vertically
averaged water abundance of 10−6 per H2. The water vapor peak occurs at a radius significantly more
distant than that expected for a passive continuous disk around a 0.6 M⊙ star, representing a volatile
distribution in the TW Hya disk that bears strong similarities to that of the solar system. This is
observational evidence for a snow line that moves outward with time in passive disks, with a dry inner
disk that results either from gas giant formation or gas dissipation and a significant ice reservoir at
large radii. The amount of water present near the snow line is sufficient to potentially catalyze the
(further) formation of planetesimals and planets at distances beyond a few AU.
Subject headings: planetary systems: protoplanetary disks -- astrochemistry -- stars: individual (TW
Hya)
1. INTRODUCTION
Giant planets and planetesimals form in the chemi-
cally and dynamically active environments in the inner
zone of gas-rich protoplanetary disks (Pollack et al. 1996;
Armitage 2011). Processes taking place during this crit-
ical stage in planetary evolution determine many of the
parameters of the final planetary system, including the
mass distribution of giant planets, as well as the chemical
composition of terrestrial planets and moons.
Indeed, some of the most important features of mature
planetary systems may be dictated by the behavior of the
water content of protoplanetary disks. Such processes in-
clude the formation of a strong radial dependence of the
ice abundance in planetesimals -- the snow line (Hayashi
1981). The deficit of condensible water inside the snow
line leads to a need for radial mixing of icy bodies during
the later gas-poor stages in the evolution of the plan-
etary system, in particular to explain the existence of
water on the Earth (Raymond et al. 2004). Ice mantles
on dust beyond the snow line allow grains to stick at
higher collisional velocities allowing for efficient coagu-
lation well beyond the limits of refractory grain growth
alone (Blum & Wurm 2008). The combined mass of the
solid component in protoplanetary disks is likely dom-
inated by volatile, but condensible, species, the most
[email protected]
1 Division of Physics, Mathematics & Astronomy, MC 150-21,
California Institute of Technology, Pasadena, CA 91125, USA
2 Space Telescope Science Institute, Baltimore, MD 21218,
USA
3 National Optical Astronomy Observatory, 950 N. Cherry
Ave., Tucson, AZ 85719, USA
4 Division of Geological & Planetary Sciences, MC 150-21,
California Institute of Technology, Pasadena, CA 91125, USA
abundant of which is water (Lecar et al. 2006).
Thus, core accretion models that rely on the avail-
ability of
large concentrations of solid material sug-
gest that giant planets will initially be found beyond
the snow line, as their formation requires the existence
of a massive ice reservoir (Kennedy & Kenyon 2008;
Dodson-Robinson et al. 2009). The added solid den-
sity offered by condensible water may also offer a way
out of the "meter-barrier" for planetesimal formation
(Weidenschilling 1997),
in which centimeter to meter
size icy bodies migrate inwards and are lost to the cen-
tral star on time scales much shorter than those of
steady state grain aggregation (Birnstiel et al. 2009): Be-
yond a certain mass density of solids, local concentra-
tions of boulder-sized particles may overcome the effects
of turbulence acting to disperse them and will grav-
itationally contract to form Ceres-mass planetesimals
(Johansen et al. 2007, 2009).
Given the central role that water plays in the formation
of planetary systems, its distribution and abundance in
protoplanetary disks has been the subject of intense the-
oretical study (e.g., Ciesla & Cuzzi 2006; Garaud & Lin
2007; Kretke & Lin 2007). Yet, the only source of obser-
vational constraints on the initial radial distribution of
water in planet-forming regions has been the distribution
of ice in the current Solar System (Hayashi 1981), the re-
maining bodies of which represent only a small part of
the story. Apart from the fact that the solar system wa-
ter distribution may be unique, theoretical study shows
that the initial distribution of ices in protoplanetary disks
evolves as does the disk temperature distribution (e.g.
Garaud & Lin 2007), and that in later stages it is con-
founded by the dynamical interactions between planets
2
and the planetesimal swarm (Gomes et al. 2005). The fi-
nal distribution of water in a planetary system is thus an
obscured tracer of initial conditions; and to understand
the role of water and other volatiles in the formation
of planets, we should measure their distribution before,
during and after planet(esimal) formation.
The most direct way to trace the water abundance dis-
tribution in protoplanetary disks is to image line emis-
sion from water vapor. However, due to the strong radial
gradient of gas temperature in disks, each line of a given
excitation energy will only trace a limited areal extent of
the disk. For instance, high-J rovibrational transitions in
the mid-infrared waveband are sensitive to the warmer
(& 300 K) and inner parts of the disk surface (. 10 AU).
Far-infrared transitions arise from cooler gas (. 300 K)
in the outer (& 10 AU) and deeperregions of the disk. In
order to measure the total water vapor content in the
disk surface, it is necessary to observe lines across states
of widely varying excitation energies.
That line emission from water vapor is common in
disks is now well established thanks to observations with
the Spitzer IRS instrument, which has detected a forest
of emission lines in the 10 -- 35 µm range. Carr & Najita
(2008) and Salyk et al. (2008) first reported a large num-
ber of rotational lines due to warm water in AA Tau, AS
205N, and DR Tau, while Pontoppidan et al. (2010b) re-
port a water emission detection rate of ∼50% in their
Spitzer survey of 46 disks surrounding late-type (M-G)
stars. With upper energy levels of Eup ∼ 1000 − 3000 K,
these water lines arise from the inner few AU region of the
disks. Detailed radiative transfer models show that the
surface water vapor abundance may need to be truncated
beyond ∼1 AU for a typical classical T Tauri star (cTTs)
disk in order to match the general line ratios over 10 --
35 µm (Meijerink et al. 2009). Follow-up ground-based
spectra of the brightest targets have constrained the gas
kinematics, and confirmed that the water vapor resides
inside the snow-line (Pontoppidan et al. 2010a).
Recently, complementary spectral tracers of water va-
por in the outer disk have become accessible via Her-
schel Space Observatory PACS and HIFI observations.
Riviere-Marichalar et al. (2012) describe the discovery of
warm water emission at 63.3 µm in 8 out of their 68
T Tauri disk sources, and tentative detections of wa-
ter transitions with similar excitation energies have also
been reported for the Herbig Ae/Be star HD 163296
(Meeus et al. 2012; Fedele et al. 2012). The first sen-
sitive search for ground-state emission lines of cold wa-
ter vapor in DM Tau indicated that the vertically av-
eraged water vapor abundance in the outer disk of this
source is extremely low, < 10−10 per hydrogen molecule
(Bergin et al. 2010). Finally, two ground state water
emission lines in TW Hya were detected with HIFI
(Hogerheijde et al. 2011).Again, the vertically averaged
∼ 10−10 per hydrogen, while
water vapor abundance is <
the estimated peak water vapor abundance is closer to
∼10−8−10−7 in the near surface layers of the disk. These
are values that can be produced by UV photodesorption
from icy dust grains. In this interpretation, the low ex-
citation water lines trace a large, but otherwise unseen,
reservoir of ice in the outer disk.
Here we present an observationally-based method for
reconstructing the water vapor column density profile
in the surface of a protoplanetary disk, based on multi-
wavelength, multi-instrument IR/submillimeter spectra,
and apply the method to the TW Hya transitional disk.
We chose TW Hya because of the availability of deep
archival Spitzer and Herschel water spectroscopy and
the existence of detailed structural models. Further, we
report on the spectroscopic identification of warm water
vapor emission at 20 -- 35 µm, a detection that permits, in
combination with the Herschel data, the first localization
of the snow line in a transitional disk.
The outline of the paper is as follows: In Section 2,
we describe the Spitzer and Herschel data.
In §3, we
describe the structural gas/dust model for the TW Hya
disk. Section 4 applies a two-dimensional line radiative
transfer model to the model disk structure to retrieve a
radial abundance water profile based on the full spectro-
scopic dataset. The results are discussed in §5.
2. DATA REDUCTION
The Spitzer IRS 10−35 µm high resolution mode
(R∼600) spectra of TW Hya were acquired as part of
a survey program of transitional disks (PID 30300), with
J. Najita as PI. The observations were done in two
epochs using different background observation strate-
gies, for an observation log see Najita et al. (2010). In
the first epoch, TW Hya was observed on source (AOR
18017792), followed by north and south off-set observa-
tions (AORs 18018048, 18018304). In the second epoch
(AOR 24402944), TW Hya was observed using a fixed
cluster-offsets mode, such that the background scans
were observed in the same sequence. All of the datasets
were extracted from the Spitzer archive and reduced us-
ing the Caltech High-resolution IRS Pipeline (CHIP) de-
scribed in Pontoppidan et al. (2010b), which takes full
advantage of the existence of redundant background ob-
servations. CHIP implements a data reduction scheme
similar to that developed by Carr & Najita (2008).
In an extensive analysis of the 10-20 µm Short-Hi (SH)
spectrum, Najita et al. (2010) report the detection of ex-
cited OH emission lines and a host of other molecular
features from TW Hya, but no water emission above a
lower flux limit of ∼10 mJy. As stressed by Najita et al.
(2010), the lack of detectable water emission from the
highly excited water lines below 20 µm does not preclude
the presence of cooler water vapor, and in a follow-up pa-
per (in prep), these authors interpret the SH and Long-Hi
(LH) data together. Our independent reduction of the
deep Spitzer IRS observations at longer wavelengths dis-
plays numerous water emission lines in the 20−35 µm
LH module. The strongest transition at 33 µm can be
clearly seen in all LH epochs. The spectrum from the
fixed cluster-offsets AOR has the highest signal-to-noise
ratio (SNR), and our analysis is therefore based on this
spectrum (presented in Figure 1). The SH spectrum is
scaled by a factor of 1.25 to match the flux of LH spec-
trum and IRAS 12 µm photometry data.
The Herschel HIFI line fluxes for the ground state
ortho and para water
transitions are taken from
Hogerheijde et al. (2011). Somewhat higher excitation
lines are probed by the PACS instrument, and narrow-
range line spectra of TW Hya from 63−180 µm were ac-
quired as part of the public Herschel Science Demonstra-
tion Phase program (observation ID 1342187238). The
PACS data were reduced using the Herschel interactive
processing environment (HIPE v.6.0), up to level 2. After
level 1 processing, spectra for the two nod positions were
extracted separately from the central spaxel. The spectra
of the two nod positions were uniformly rebinned, using
an over-sampling factor of 2 and an up-sampling factor
of 2, before co-adding to produce the final result. Linear
baselines were fitted to the local continua and integrated
line fluxes (see Table 2) were obtained using Gaussian
fits. While no water emission is formally detected with
PACS at 3σ, there are indications of water lines consis-
tent with our water abundance model predictions; see
Section 4 for further discussion.
3. THE TW HYA DISK STRUCTURE
The accuracy of retrieved chemical abundances in pro-
toplanetary disks is coupled to how well the two- or
three-dimensional gas kinetic temperature and volume
density structures are known. Our model development
consists of three stages: First, we construct an axisym-
metric model of the dust component of the disk, based on
fits to the continuum spectral energy distribution (SED),
using the RADMC code(Dullemond & Dominik 2004).
We then assume that the gas density distribution fol-
lows that of the small dust grain population -- a rea-
sonable assumption since the small grains carry most of
the disk opacity and are dynamically well-coupled to the
disk gas. However, we do allow thermal decoupling of the
gas from the dust near the disk surface, in the form of
a constant temperature offset constrained using the CO
rovibrational emission spectrum near 4.7 µm. Finally, we
constrain the water vapor radial column density distribu-
tion (and thus abundance) at all disk radii via fits to the
multi-wavelength water spectra acquired for TW Hya.
3.1. The dust temperature and density structure
Dust, as the dominant source of disk continuum opac-
ity, determines the disk SED shape. At a distance of
only ∼ 51 ± 4 pc (Mamajek 2005), TW Hya is one of
the best studied protoplanetary disks, with extensive
photometric data available from UV to cm wavelengths.
Calvet et al. (2002) developed a physically self-consistent
dust structure model for TW Hya. They found that
the disk is vertically optically thin within 4 AU, corre-
sponding to a depletion of small dust grains by orders
of magnitude at these radii. This inner zone of dust de-
pletion has been confirmed by subsequent observations
(Eisner et al. 2006; Hughes et al. 2007). A small amount
of dust in the inner disk is needed to explain the 10 --
25 µm amorphous silicate features (Calvet et al. 2002),
and accretion onto the star continues. The existence of
hot gas in the innermost disk has been further confirmed
by the detection of CO ∆v=1 rovibrational emission
near 4.7 µm (Rettig et al. 2004; Salyk et al. 2007, 2009)
and via Spectro-Astrometric (SA) observations of these
same lines (Pontoppidan et al. 2008). It is not yet clear
whether the gas content has been depleted to the same
degree as the dust in the inner disk, although Gorti et al.
(2011) estimated enhancements of 5-50 in gas/dust in in
the inner disk.
We adopt M⋆ = 0.6 M⊙, R = 1 M⊙ and Teff = 4000 K
for the stellar mass, stellar radius and effective temper-
ature (Webb et al. 1999). The inclination angle of the
disk axis to the line-of-sight is fixed to i = 7◦ (Qi et al.
3
Table 1
Disk parameters for the dust model
Parameter
Inner disk Cavity wall Outer disk
r (AU)
p
γ
0.055 − 4
0.8
0.3
4 −5
0.0
-0.25
5−196
1.0
0.2
Note. -- Here γ is the index that describes the
vertical scale height zd: zd = z0(r/r0)1+γ where z0 =
0.7 at r0 = 10 AU; p is the index of the surface density
distribution: Σd = Σ0(r/r0)−p.
2004). The standard gas-to-dust ratio of 100 is used to
estimate the total disk mass.
The dust structure model adopted is similar to
Andrews et al. (2012): an optically thin inner disk be-
tween the sublimation radius (rsub) and the cavity edge
(rcav), a cavity "wall" and an optically thick disk. The
cavity "wall" component is needed to reproduce the spec-
tral shape of TW Hya from 10−30 µm (Uchida et al.
2004).
The sublimation radius was set to rsub =
0.055 AU, the location where dust temperature reaches
1400 K. The cavity radius is fixed as rcav = 4 AU, in ac-
cord with previous near-IR interferometry (Eisner et al.
2006; Hughes et al. 2007; Akeson et al. 2011). We adopt
the power-law density profile of Andrews et al. (2012) for
the outer disk, which is based on 870 µm continuum in-
terferometric imaging. A summary of the dust structure
model parameters can be seen in Table 1.
Following Pollack et al. (1994) and D'Alessio et al.
(2001), we use a dust mixture of astronomical silicates,
organics and water ice. For the optically thin inner disk,
pure glassy silicate is used to match the strong silicate
bands at 10 and 18 µm. The dust size distribution is
taken to be n(a) ∝ a−s with s = 3.5 between amin = 0.9
and amax = 2.0 µm (Calvet et al. 2002). In the cavity
wall, we use a dust size range between 0.005 and 1 µm;
outside the cavity wall, 95% (by mass) of the dust has a
size distribution extending from 0.005 µm to 1 mm, with
the remaining 5% contained in grains from 0.005 -- 1 µm.
Total dust masses in the optically thin and thick parts
of the disk are 2.77 × 10−9 and 4.4 × 10−4 M⊙. The SED
fit and our fiducial gas temperature and density profiles,
discussed further below, are presented in Figure 2.
3.2. The gas density and temperature
The gas/dust ratio in disks evolves due to grain growth
and transport processes (Birnstiel et al. 2009), and likely
has significant radial and vertical structure. For simplic-
ity, the gas density is here assumed to follow that of the
dust, with a constant gas/dust mass ratio throughout
the disk. For TW Hya, the estimated global gas/dust
ratio varies from 2.6 to 100 (Thi et al. 2010; Gorti et al.
2011). The recent detection of the HD J = 1-0 line from
TW Hya has offered an independent and robust estima-
tion of the gas mass (Bergin et al. 2013), which indicates
a globally averaged gas/dust ratio close to 100, a value
we adopt here. Significant dust vertical settling will pro-
duce an enhanced gas-to-dust ratio in the upper layers
of the disk. Such settling should have little impact on
the longest wavelength water transitions studied here,
or on the strongest, optically thick lines traced by the
Spitzer IRS. By creating a larger column of gas above
the τdust = 1 surface, the absolute fractional abundance
4
Figure 1. The Spitzer IRS high resolution (R=600) spectrum of TW Hya from 10-35 µm; the trace at bottom presents the data after the
subtraction of a spline continuum fit (and scaled by a factor of 3 to show the emission lines more clearly). No water emission is detected
in the 10-20 µm SH modules (Najita et al. 2010), but the 20-35 µm LH spectrum shows clear evidence for water vapor. The short vertical
ticks indicate water vapor rest wavelengths (HITRAN 2008 database) for lines with intensities >5×10−22 cm−1/(molecule cm−2).
Figure 2. Left panel: A 2-D model fit to the TW Hya SED using RADMC and the parameters in Table 1. The stellar spectrum (peaking
near 1 µm) is generated from a Kurucz model with Teff = 4000 K. The photometric points are taken from Rucinski & Krautter (1983),
2MASS, Low et al. (2005), and Weintraub et al. (1989), while the 5 -- 35 µm Spitzer IRS spectrum is that displayed in Figure 1. Middle
panel: Fiducial gas temperature profile. Right panel: Fiducial gas density profile, with the Av=1 height depicted by the solid line.
of water derived by a well mixed LTE model likely pro-
vides an upper bound, but the inferred radial structure
in the water vapor column density should be fairly robust
against changes to the dust distribution.
Above a certain altitude in the disk, where the en-
vironment becomes exposed to the ambient radiation
field, the gas can also be thermally decoupled from the
dust. The gas will adopt a temperature profile that is
a balance among heating processes, such as mechanical
heating via accretion or that driven by photodissocia-
tion or the photoelectric electric effect, and cooling rates
driven by atomic, molecular, and dust grain emission
(Glassgold et al. 2004; Kamp & Dullemond 2004). Gen-
eral gas temperature profiles can be estimated using de-
tailed thermo-chemical models (e.g. Woitke et al. 2009;
Najita et al. 2011). However, such profiles can be highly
dependent on input assumptions, and interdependencies
between model parameters and observables can be ob-
scure. Here, we simplify the process by deriving the verti-
cal gas temperature structure in the inner disks using the
rotational ladder from M-band CO P-branch v = 1 − 0
emission lines. Due to its (photo)chemical robustness,
the abundance of CO is predicted to be fairly constant
at nCO/nH2 ≈ 1.2 × 10−4 in regions warmer than 20 K
(Aikawa et al. 1996). A high CO abundance is expected
to persist even for a depleted inner disk, such as that of
TW Hya. Indeed, in the chemical model of Najita et al.
(2011), the abundance of CO rises to ∼ 10−4 once the
vertical column density of H2 reaches 1021 cm−2 for radii
beyond 0.25 AU (physical densities are >109 cm−3).
Since the heating of the inner disk gas is driven by
X-ray and FUV photons from the central star and stel-
lar accretion flow (Kamp & Dullemond 2004; Gorti et al.
2011), we assume that gas and dust temperature become
decoupled in regions where the radial optical depth for
visible photons is less than unity. That is:
Tgas =(cid:26) Td,
Av > 1
Td + δT, Av 6 1 ,
(1)
5
gas density structure follows that of the dust.
We stress that such gas/dust thermal decoupling
should have only a modest impact on a principle molec-
ular mapping result presented here, namely the signifi-
cant drop in the water vapor column density beyond the
snow line. For the highest excitation lines measured by
Herschel (and Spitzer ) that trace the photon-dominated,
and thus heated, layers in the inner disk and cavity wall,
gas densities are at least ∼109 H2/cm3. Under such con-
ditions the simulations of Meijerink et al. (2009) show
that the emergent fluxes of the water lines detected here
are within factors of two-three of their LTE values. Thus,
the water vapor column densities should be reasonably
well determined by LTE calculations whose temperature
distributions are constrained by the CO M-band obser-
vations.
The outer disk is too cold to emit in such high excita-
tion water lines, and as discussed further in Section 5, the
physical density in the Av∼1 layer at radii beyond 5-10
AU is only ∼107 H2/cm3, some 100-1000× less than the
critical densities of the mid- to far-infrared water vapor
lines examined here. Only the lowest-excitation water
transitions measured by Herschel principally trace the
outer, deeper disk layers. Here the M-band CO obser-
vations are of little use, but the lowest-J pure rotational
lines do probe the relevant disk gas. In RADLite sim-
ulations of lines up to J = 6-5 we find, in agreement
with Qi et al. (2004), that significant CO depletion in
the cold disk midplane is needed to match the observed
line fluxes. Water can be expected to deplete in a simi-
lar manner, with the bulk of the vapor emission arising
from an intermediate layer at densities where LTE cal-
culations offer column density estimates good to within
an order-of-magnitude (Hogerheijde et al. 2011).
4. RETRIEVING THE RADIAL WATER VAPOR PROFILE
4.1. The water line model
Given the gas temperature and density structure for
TW Hya, we can now constrain the water vapor con-
tent of the TW Hya disk surface as a function of radius.
Our goal is not to provide an exacting description of the
volatile abundances versus radius and height, but to de-
termine what radial distribution of water vapor is most
consistent with the available data.To this end, we con-
struct a step function in water vapor abundance, with
one value, Xinner, in the inner gas-depleted disk within
4 AU, another, Xring, in a ring starting at 4 AU and ex-
tending to the snow line at Rsnow, and an outer disk
abundance Xouter. That is,
XH2O =( Xinner, R < 4 AU
Xring,
Xouter, R > Rsnow
4 AU 6 R 6 Rsnow
(2)
As described below, our calculations assume LTE but do
account for line and continuum opacity. The XH2O val-
ues derived thus reflect the water vapor column densities
to different depths into the disk.
Inside of 4 AU and
for the longest wavelength Herschel lines sensitive to the
outermost radii the dust opacity permits the full vertical
extent of the disk to be sampled. For the IRS features
and the shorter wavelength PACS lines that sample gas
near the transition radius, significant dust opacity limits
the fitted column densities, and hence the XH2O values,
Figure 3. Model CO rovibrational fluxes from decoupled gas/dust
temperature models compared with the observational data (de-
picted as crosses with error bars, from Salyk et al. 2007). The
models show three different values of δT, the temperature differ-
ence between gas and dust, with values of 120 (diamonds), 135
(squares) and 150 K (circles).
where Td is dust temperature in the disk at (r, θ) in
spherical coordinates, Av is the visual extinction along
the radial direction θ, and δT is a free parameter that
describes the gas-dust temperature difference, to be de-
termined through fits to observational data.
The v=1-0 P(1-12) CO line fluxes at 4.7 µm are
taken from Salyk et al. (2007), while the model spec-
tra are generated by the raytracing code RADLite
(Pontoppidan et al. 2009). The lines are ∼7.5 km/s wide
(FWHM) (Pontoppidan et al. 2008), consistent with the
nearly face-on orientation of TW Hya. The inner edge
of the CO emission in the inner disk is set at rin =
0.11 AU, based on SA imaging of the 4.7 µm line emis-
sion (Pontoppidan et al. 2008). The outer edge of the
CO-emitting zone is not well constrained, and is set to
4 AU, the dust transition radius. The model is relatively
insensitive to the size of the outer edge, since the major-
ity of the emission is produced at radii ≪ 4 AU.
The best fit has δT=135±15 K (Figure 3). From M-
band rotation diagram fits, Salyk et al. (2007) derive
NCO = 2.2 × 1018 cm−2, corresponding to a vertical
total gas column density of NH∼1022 cm−2. The 135 K
gas-to-dust temperature difference is consistent with the
detailed gas temperature balance model of Najita et al.
(2011), whose treatment yields δT∼100 K at r=0.25 AU
for an integrated column density of NH = 1022 cm−2.
RADLite simulations predict a line width (prior to in-
strument profile convolution) of FWHM∼6.8-9.6 km/s
for vturb ∼ 0.05 vKepler. This is somewhat larger than
observed 7.5 km/s linewidths, but this difference may be
accounted for by the difference in disk inclination relative
to that (i = 4◦) derived by Pontoppidan et al. (2008).
The best fit δT of 135 K applies only to the inner disk
radii probed by CO, but similar physics will decouple the
gas and dust temperatures near the disk surface at larger
radial distances. To model this decoupling, we adopt
Tg = Td +δT ×e−r/50 AU for the gas temperature at large
scale heights, a parameterization that matches broadly
the models of Thi et al. (2010) for TW Hya. Again, the
6
to the upper layers of the disk. Freeze-out and dust/line
opacity greatly limit access to the midplane beyond 4
AU.
Given the wide span in upper level energies of the lines
considered, it is generally possible to identify subset of
lines that co-vary. Consequently, we can explore one
model parameter at a time to derive a best-fit model, il-
lustrated in Figure 4. As we shall see, the simplest LTE
model, namely a constant water abundance throughout
the disk, is inconsistent with the data in hand.In Sec-
tion 4.2.2, we further discuss whether the data support
a modification to this basic structure in which the drop
in water abundance beyond the snow line instead occurs
over some finite, measurable distance.
To render model spectra, we set the level populations
to LTE. While the critical densities (ncrit = 1010 − 1012
cm−3) of the high energy mid-IR water transitions sug-
gest that a non-LTE treatment is needed, there is cur-
rently no robust and fully tested non-LTE framework for
the modeling of infrared water lines. That is, a non-
LTE calculation is also likely to be inaccurate, given
the uncertainty of collisional rates and the complexity
of the transition network. Further, previous work sug-
gests that non-LTE effects may alter line fluxes by factors
of only a few for the moderate excitation lines detected
here (Meijerink et al. 2009; Banzatti et al. 2012), which
will preserve the qualitative aspects of our treatment. In
setting level populations to LTE, we make it easier to
reproduce our results and to evaluate the validity of our
retrievals using more detailed models in the future.
We assume a subsonic turbulent velocity broaden-
ing of vturb ∼ 0.05 vKepler, and a constant ortho-to-
para (O/P) water ratio of 3:1, that consistent with the
mid- and far-infrared lines (Pontoppidan et al. 2010b).
The cold gas seen by HIFI has a lower O/P ratio of
0.7 (Hogerheijde et al. 2011), but the difference does
not substantially affect the derived abundance struc-
ture. Model spectra are typically rendered with a grid
of 0.3 km/s, and then convolved with a Gaussian instru-
ment response function to match the observed spectra.
The Herschel HIFI spectra are rendered with a finer ve-
locity grid of 0.05 km/s.
4.2. Best-fitting model
The best-fitting radial water column density profile in
TW Hya is shown in Figure 4, whose spectral predictions
are compared to the Spitzer LH data in Figure 5. We
find that the inner optically thin region of TW Hya is
dry, with upper limits on the vertically integrated frac-
tional water abundance of Xinner < 10−6. The water de-
tected by Spitzer originates in a thin ring near the disk
surface, starting at 4 AU and ending at a snow line just
beyond this at Rsnow ∼ 4.2 AU. At larger radii, the frac-
tional water abundance drops abruptly by several orders
of magnitude, to Xouter values near 3.5×10−11.
The uniqueness of this solution is illustrated in Figure
6, which shows how variations in the water abundance
in different regions of the disk affect different lines. This
differential sensitivity is the foundation of our method
for retrieving the radial column density, and thus abun-
dance, profile, and is the subject of the following discus-
sion.
Figure 4. The best-fit radial abundance of water vapor in TW
Hya, illustrated by the vertically integrated column density profile
(solid curve). Within 4 AU, the maximum water abundance is
depicted by arrows. The dashed curve is the total gas column
density. At the top of the figure, the role of multi-wavelength
observations in tracing different radii of the disk are highlighted,
along with representative facilities of each wavelength window.
Figure 5. Detailed comparisons of the Spitzer IRS spectrum
(black) with RADLite LTE water models (red). The spectrum of
RNO90, one of the strongest water emitters in Pontoppidan et al.
(2010b), is plotted at bottom, to guide the eye. The vertical dashed
markers show the location of OH lines not discussed in this paper.
4.2.1. The inner disk
The inner disk of TW Hya -- radii within 4 AU -- is par-
tially cleared out, leading to integrated vertical column
densities that are more than two orders of magnitudes
lower than those typical for cTTs. The non-detection
of water lines between 10 and 20 µm, in the Spitzer SH
range, may be a reflection of this, but even though the
column density in the inner disk is dramatically reduced
from the extrapolation of the outer disk to smaller radii,
the scale height is sufficiently small that the physical den-
sities at the Av∼1 surface exceed 109 H2/cm3. Thus,
non-LTE effects are likely to be fairly unimportant, and
we are able to produce meaningful upper limits on the
fractional water abundance at r <4 AU.
We investigated the effects on the Spitzer and Herschel
lines when varying the inner disk abundance ratios be-
tween Xinner = 10−7 and 10−4.
Interestingly, we find
Observed and calculated water line fluxes
Table 2
Type
Transition
λ( µm)
Eup (K)
A coefficient(s−1) Flux(10−19W/m2) Model(10−19W/m2)
7
o-H2O 1139-10010
p-H2O 1129-10110
o-H2O 836-707
p-H2O 981-872
o-H2O 982-871
o-H2O 845-716
o-H2O 1166-1055
p-H2O 1156-1047
p-H2O 853-744
p-H2O 761-652
o-H2O 762-651
o-H2O 854-743
o-H2O 634-505
o-H2O 818-707
o-H2O 707-616
p-H2O 817-808
o-H2O 423-312
p-H2O 615-524
p-H2O 322-211
p-H2O 413-322
o-H2O 212-101
p-H2O 111-000
o-H2O 110-101
17.23
17.36
23.82
23.86
23.86
23.90
23.93
23.94
30.47
30.53
30.53
30.87
30.90
63.32
71.95
72.03
78.74
78.93
89.99
144.52
179.53
269.27
539.29
2438.8603
2432.5234
1447.5970
2891.7024
2891.7022
1615.3501
3082.7639
2876.1493
1807.0023
1749.8575
1749.8506
1805.9308
933.7488
1070.7757
843.5459
1270.3910
432.1913
781.1873
296.8471
396.4126
114.3874
53.4366
60.9686
9.619E-1
9.493E-1
6.065E-1
3.128E+1
3.128E+1
1.034
1.612E+1
9.754
8.923
1.355E+1
1.355E+1
8.687
3.504E-1
1.730
1.147
3.099E-1
4.852E-1
4.501E-1
3.541E-1
3.345E-2
5.617E-2
1.852E-2
3.477E-3
< 163.7
< 617.3
4669±364a
)3750±317a
(cid:27)1883±287a
< 123.2
< 127.2
< 35.4
< 39.1
< 54.3
< 60.9
< 5.9
< 8.6
3.07± 0.19
3.46± 0.15
37.6
110.4
3051
4940
3393
84.8
77.0
11.6
76.3
33.7
45.8
8.5
10.7
3.24
1.89
Note. -- a. Water lines are often blended in the Spitzer IRS LH spectrum due to the limited resolution
of R∼600. The fluxes are measured by integrating the spectrum over each distinctive feature/region after a
linear continuum subtraction.
that inner disk fractional abundances higher than 10−6
significantly overpredict several water lines in the PACS
range, especially the high excitation 63.3 µm lines, while
the Spitzer SH lines allow somewhat larger abundances.
That is, the Spitzer SH spectrum tends to be consistent
with a model in which the lack of lines below 20 µm is
explained by the overall low gas mass in the inner disk,
and not necessarily a drier disk. It is the addition of the
non-detection of PACS lines that require the fractional
water abundance to be suppressed in the inner disk.
4.2.2. The transition region and the snow line
Next, we consider the origin of the water lines seen at
20-35 µm. Figure 6 shows that the Spitzer LH lines de-
tected are uniquely sensitive to variations in the water va-
por content around the transition region. It is clear that
a high water vapor column density is required to fit the
Spitzer spectra, with an emitting area close to π×1 AU2.
Our fiducial dust/gas model yields Xring ∼ 2.5 × 10−4
above the τdust=1 surface (line opacity is also signifi-
cant). As discussed in §5.2, the required column den-
sity/abundance can be traded off against excitation tem-
perature to some extent, with higher excitation temper-
atures corresponding to slightly smaller radii.
In the context of our transitional disk model, the most
likely location for warm water vapor lies at the transi-
tion radius, or 4 AU, a prediction that can be tested
by measurements of the water (or perhaps OH) emis-
sion line profiles using high resolution thermal-infrared
spectroscopy. Interestingly, the cavity "wall" structure
needed to reproduce the 10-30 µm SED (with a z/r value
of 0.3) has a surface area close to that derived from the
water emission for an inclination of seven degrees, further
reinforcing the likely radial location of the high water va-
por abundance.
Further, the PACS and HIFI lines are sensitive to the
location of the snow line (the outer edge of the water-
rich ring). Consequently, we varied the location of the
snow lines between 4.2-6 AU. In Figure 6, it is seen that
placing the snow line farther out in the disk over-predicts
primarily the PACS lines. Hence, the detection of water
lines by Spitzer in combination with the non-detection
of water lines in PACS waveband provides strong con-
straints on the location of the surface snow line.
We explored one more aspect of the water abundance
profile around the snow line. Because the water conden-
sation temperature is a function of pressure and because
of the strong vertical gradients in the disk, the local wa-
ter vapor abundance need not precisely be a step func-
tion in radius. In particular, can the water lines between
40-150 µm be used to constrain how rapidly the water
abundance, as inferred from the column density struc-
ture, drops beyond the snow line with radius?
Specifically, we modify the abundance step function
Reff
with an exponential drop-off beyond the snow line:
XH2O(R > Rsnow) = (Xring −Xouter)e(cid:16) Rsnow−R
(cid:17) +Xouter,
(3)
where Reff is the scale length of the abundance change.
We consider Reff = 0.1, 0.5, and 1 AU, and find that
models with Reff & 0.5 AU produce more flux than is
observed by the Herschel PACS instrument (Figure 7).
This supports our original assumption of a step function,
and we conclude that the surface snow line in TW Hya
is radially narrow; that is, it occurs over a region that is
a small fraction of its distance to the star.
4.2.3. The outer disk
An extremely low vertically averaged abundance be-
yond the snow line, Xouter=3.5×10−11, is required by
LTE fits to the HIFI ground state water lines at 267
and 537 µm.
Indeed, the strongest constraint on the
outer disk water vapor abundance is provided by the
ground-state lines, although significant limits are also
8
)
2
-
m
c
(
g
o
L
y
t
i
s
n
e
D
n
m
u
o
C
l
t
r
e
a
W
24
22
20
18
16
14
24
22
20
18
16
14
24
22
20
18
16
14
24
22
20
18
16
14
IRS SH
IRS LH
IRS LH
PACS
PACS
HIFI
a1
a2
a3
a4
a5
a6
b1
b2
b3
b4
b5
b6
c1
c2
c3
c4
c5
c6
a
1e-5
1e-6
1e-7
b
2.5e-3
2.5e-4
2.5e-5
c
3.5e-10
3.5e-11
3.5e-12
d
d1
d2
d3
d4
d5
d6
p-H2O
111-000
269µm
0.7
0.5
0.3
0.1
0
0
0.7
0.5
0.3
0.1
0
0
0.7
0.5
0.3
0.1
0
0
0.7
0.5
0.3
0.1
0
0
N
o
r
m
a
l
i
z
e
d
F
u
x
l
2
4
r(AU)
6
8
17.2017.30
λ(µm)
23.7
24.0
λ(µm)
30.5 30.8
λ(µm)
63.3
λ(µm)
63.4
179.4179.8
λ(µm)
0 1 2 3 4 5 6
v (km/s)
Figure 6. The sensitivity of selected water line fluxes to variations in the water vapor column density distribution. The first column
shows the model water vapor distributions. The best-fit model is the solid red curve, while the green and blue curves are models calculated
to illustrate the sensitivity of the various lines to changes in the water abundance at different disk radii. The numbers in the first column
indicate the fractional (per hydrogen molecule) water abundance in (a) the optically thin inner disk, (b) the ring at the transition radius
and (c) the outer disk. The final (d) panel shows the sensitivity of the model to the radius of the snow line. The following six columns
show model line spectra compared to the observed spectra for TW Hya. The ladder crossed line in the 30 µm column is OH, which is not
discussed in this paper. Please see Fig. 5 for additional OH line identifications.
provided by slightly higher-lying transitions in the PACS
range, in particular the 179.5 µm transition. The low
water vapor content of the outer disk is consistent with
the analysis of Hogerheijde et al. (2011) and the HIFI
non-detection of water lines in another transitional disk,
DM Tau (Bergin et al. 2010). Specifically, our Xouter
is comparable to the global disk-averaged value derived
by Hogerheijde et al. (2011) -- 7.3×1021 g of water in a
1.9×10−2 M⊙ disk, or XH2O=2.2×10−11 -- but as these
authors stress there is likely significant vertical structure
in the water vapor abundance and a large ice reservoir
beyond 4-5 AU.
5. DISCUSSION
From an analysis of the water emission lines from the
TW Hya disk over the 10 − 567 µm interval, both warm
(∼220 K) and cold water vapor are found to be present.
The warm water emission most likely originates in a nar-
row ring region between 4 -- 4.2 AU where abundant water
vapor carries much of the cosmically available oxygen,
XH2O ∼ 10−4. Outside of 4.2 AU, the vertically aver-
aged water vapor column density decreases dramatically
over a radial distance less than 0.5 AU, marking the loca-
tion of the disk's surface snow line. Each of the datasets
included in this study uniquely constrain one aspect of
the distribution profile, and so we demonstrate the im-
portance of the use of multi-wavelength datasets. Here
we discuss some implications of our results.
5.1. The origin of the surface water vapor
How does this study inform the origin of the observed
water vapor, and of water in protoplanetary disks in gen-
eral? There are two potential sources of water: One is in
situ gas-phase formation, while the other is grain surface
formation in a cold reservoir -- which could be the disk
itself or the primordial protostellar cloud. In the second
case, a mechanism is needed to transport the ice inwards
and upwards to a location where it can either thermally
evaporate or be photodesorbed and observed in the form
of vapor (see, for example, Supulver & Lin 2000).
9
Figure 7. Constraints on the sharpness of the snow-line in TW Hya. The left panel shows model water vapor abundance decreases from
XH2O ∼ 2.5 × 10−4 to 3.5 × 10−11 with three different radial scale lengths: 0.1 AU (dot line), 0.5 AU (solid line) and 1 AU (dash line). The
right panel shows the modeled LTE line fluxes over-plotted on Herschel PACS spectra. The line styles are the same as in the left panel.
In the warm inner disk (r<1AU, z/r<0.2), H2O
can be formed in the gas phase (Glassgold et al. 2009;
Bethell & Bergin 2009), even in the presence of sub-
stantial photolyzing radiation. The key is a sufficient
supply of H2, which drives the formation of water via
successive hydrogenation of atomic oxygen in reactions
with significant activation barriers. More precisely, we
find that using the photodissociation rate calculated only
at Lyα (which dominates the 3×10−3L⊙ FUV contin-
uum flux), and water reaction rates from Baulch, D. L.
(1972) -- kOH=3.0×10−14T e−4480/T cm3 molecule−1s−1
and kH2O=3.6×10−11T e−2590/T cm3 molecule−1s−1 --
this simplified chemical model predicts water abundances
near 10−4 when Tg &200 K, ng>109 cm−3. These condi-
tions are satisfied in the inner disk and the cavity wall,
but not necessarily in the photon-dominated upper layers
of the outer disk.
Thus, production of water via gas-phase reactions can
potentially explain the high abundance of ∼220 K water
vapor observed inside the TW Hya snow line, even if the
initial conditions are atomic, and only until the elemental
abundance of oxygen is depleted (∼ 5 × 10−4 relative
to H, Frisch & Slavin 2003).
Indeed, modern thermo-
chemical models (Woitke et al. 2009; Najita et al. 2011;
Walsh et al. 2012) predict both that much of the oxygen
is locked up in water vapor in the warm inner disk and
at large scale heights where the gas is heated by UV and
X-ray photons.
The best fit water content just inside the TW Hya
snowline is not sufficiently high to distinguish between in
situ gas-phase formation and evaporating icy bodies, but
is consistent with thermo-chemical models. Well inside
the transition radius, the water abundance drops by two
orders of magnitude, in conflict with this simple model.
However, the models by Woitke et al. and Walsh et al. do
not include the depleted inner region of TW Hya, making
direct comparisons difficult for the innermost disk.
The warm gas-phase chemistry will not operate in
the outer disk. Here, the primordial surface ice chem-
istry will dominate (Tielens & Hagen 1982), resulting
in a huge reservoir of icy dust and larger bodies from
which water molecules must be thermally evaporated
or desorbed by high-energy particles and photons. As
discussed in Bergin et al. (2010) and Hogerheijde et al.
(2011), models of photodesorption from icy grains find
that a depletion of icy grains from the outer disk sur-
face, presumably by setting to the midplane, is needed
to reproduce the low observed outer disk water vapor
abundances. A zeroth-order expectation is that the wa-
ter vapor abundance across the snow-line can be treated
as a step function (Ciesla & Cuzzi 2006), with high in-
ner disk water vapor abundances created by a mixture
of gas-phase reactions and evaporating icy bodies and a
low outer disk abundance maintained by desorption.
We find a very low (vertically averaged) outer disk wa-
ter abundance of a few times 10−11 per hydrogen, con-
sistent with previous Herschel-HIFI observations, but in
conflict with the static thermo-chemical models, which
indicate surface abundances of 10−7 − 10−9 per H. We
interpret this as further evidence for settling of icy grains
in the outer disk of TW Hya.
Another potential test for icy-grain contributions to
the observed disk water vapor content would be a mea-
surement of the O/P ratio. For gas phase formation of
water at >200 K, an O/P ratio of 3 is expected, while the
O/P ratio for evaporation or sputtering can be consider-
ably less than 3 for grain surface temperatures less than
∼ 40 K. However, for higher temperature ices formed in
regions close to the snow line, the O/P ratio may still
be near 3. Thus, detailed measurements of the O/P
ratio will be needed to trace the origin of the snow-
line water vapor, such as could be obtained, for exam-
ple, with next generation ground-based (VLT-VISIR) or
space-based (JWST-MIRI) thermal-IR spectrographs.
10
Figure 8. The χ2 surface for LTE slab model water fits for TW
Hya as a function of N and T , with an effective radius R=1.0 AU.
The contour plot is based on Spitzer IRS LH spectra, with best-fit
parameters of T = 220 K and N =1018 cm−2. The shadowed area
depicts the parameter space that is excluded by the water upper
flux limits from the Spitzer IRS SH spectrum. Empty squares
depict the best LTE slab model fit results for cTTs with water
detections in the Spitzer -IRS wavelength range (Salyk et al. 2011).
5.2. The water vapor abundance distribution in
transitional disks
The Spitzer IRS spectrum reveals that the water vapor
emission from TW Hya is different from that of typical,
but less evolved, protoplanetary disks around solar-type
stars. A comparison of the basic properties of the TW
Hya water emission is shown in Figure 8, where an LTE
slab model (disk-averaged) temperature and column den-
sity are compared to those of cTTs disks (Salyk et al.
2011). This is consistent with the detailed radial abun-
dance structure derived for TW Hya.
Although the relative contributions to the observed
water vapor in TW Hya from gas-phase reactions and
evaporating icy planetesimals is unknown, it may be the
ideal target to observe the evaporation process as its rel-
atively small 'hole' size allows the snow line to reside in
the optically thick outer disk. In contrast, many of the
known transitional disks have transition radii of 10 AU or
greater (Andrews et al. 2011), and the snow-line would
be located well inside the optically thin region of the disk.
They also typically lie two or three times more distant
than TW Hya. Therefore, comparisons with additional
transitional disks may need to wait for the availability of
more sensitive instruments.
In contrast to the ∼1 AU snow line found by
Meijerink et al. (2009) for cTTs disks, the TW Hya
snow-line is located at a significantly larger radius, in
spite of its lower luminosity as compared to DR Tau and
AS 205 N. In the case of transitional disks, however, the
exposure of the transition radius to nearly unobstructed
stellar light and the 'face on' nature of the cavity wall
at the transition radius should heat the disk more read-
ily at large radii and push the snow-line outwards. One
interesting consequence of this property is that if giant
planets form and clear out a hole in the disk, creating a
transitional disk, the subsequent migration of the snow
line to larger radii could slow the growth of new plane-
tary cores over a larger range of radii, out to the location
of the new snow line.
At the same time, new planetesimal formation may
be catalyzed near the new snow line due to enrichment
of water by a combination the radial cold finger effect
of Stevenson & Lunine (1988) and inwards radial migra-
tion of icy bodies (Ciesla & Cuzzi 2006). The observed
high abundance of water around 4 AU in TW Hya is
supportive of a scenario in which new icy planetesimal
formation, and perhaps even giant planet formation, is
ongoing at this location.
Another interesting feature of the distribution of water
vapor abundance in TW Hya is the relatively dry inner
disk. The water vapor abundance in the innermost disk
(< 0.5 AU) is actually unconstrained by the Spitzer data
due to beam dilution, but for r >0.5 AU the abundance
upper limit is two orders of magnitude below that esti-
mated in cTTs disks, and the lack of water in inner disk
can be explained by two possible scenarios.
First, the inner disk really is empty between 0.5-4 AU,
a scenario consistent with near-IR interferometric data.
Akeson et al. (2011) found an inner optically thick ring
around 0.5 AU plus an outer optically thick disk starting
at 4 AU can match the full suite of near-IR interferomet-
ric data and the general SED shape of TW Hya. Recent 8
- 18 µm speckle imaging, however, has suggested a more
continuous dust distribution out to 4 AU and perhaps
the presence of a companion (Arnold et al. 2012).
Further, analysis of the ALMA CO J=2-1 and 3-
2 Science Verification observations of TW Hya has re-
vealed gas signatures down to radii as close as 2 AU
(Rosenfeld et al. 2012). Thus, the more likely scenario
for the structure of the TW Hya disk is that outlined
in Figure 4, namely an inner disk in which CO remains
abundant since the kinetic temperature is much higher
than that needed for freeze-out (∼ 20 K).Water in the
inner disk can still be subjected to some depletion, or
in the case that a companion gates the accretion flow
into the inner disk any water-ice rich grains that settle
to the (outer disk) mid-plane may be blocked from fur-
ther inward migration. If this is the case, there should
be a large difference in the gas phase C/O ratios be-
tween the inner and outer parts of the disk. Thus, the
quantity of gas along with its C/O content are poten-
tially probes that can distinguish whether a given tran-
sition disk has been sculpted by planet formation or other
mechanisms (grain growth or photo-evaporation, for ex-
ample; see Najita et al. 2011 for a further discussion of
how changes in the C/O ratio can impact disk chem-
istry).
A final aspect of these observations is the short scale
length (<0.5 AU) over which the dramatic water vapor
decrease occurs, a distance in conflict with scenarios that
consider only freeze-out. Such a radial profile may be
further evidence for a vertical cold finger effect, in which
the surface snow-line location reflects that at the mid-
plane, perhaps due to mixing (Meijerink et al. 2009).
5.3. Molecular mapping with multi-wavelength spectra
We have demonstrated here how multi-wavelength
spectroscopic data can be used to reconstruct molecu-
lar abundances, even when the observations are spatially
(and often spectrally) unresolved.The snow line differ-
ences inferred for the tiny sample observed to date are in
accord with the changing physical conditions that pas-
sive cTTs disks experience as they evolve. Much more
detailed constraints and comparisons to models will be-
come available over the coming years by exploiting the
significant infrared database that exists for stars+disks of
varying mass and evolutionary state (Pontoppidan et al.
2010b; Salyk et al. 2011) along with new observations at
longer wavelengths.
Because protoplanetary disks are complicated struc-
tures, the multi-wavelength molecular mapping method
should be used with some caution, however. The flux
of each line will arise from a range of disk regions, with
lines at various wavelengths characterizing different radii
and/or vertical depths, and an overview of several of the
key water vapor tracers for the best fitting TW Hya
model is presented in Fig. 9. As demonstrated in this
work, it can take considerable effort to build realistic
models for an individual source, and efforts herein were
aided by a large set of ancillary observations of this well-
studied disk. Nevertheless, it is important to stress that
relatively fewer uncertainties are introduced in this type
of modeling than with the use of full thermal-chemical
disk models, in which chemical abundances are sensitive
to a wide array of model parameters, including FUV and
X-ray fluxes, accretion rate, disk viscosities and trans-
port rates, dust-versus-gas settling geometries, grain-
surface reaction rates, and so on (e.g. Heinzeller et al.
2011). Further, studies of large disk samples will require
fast, robust models. We therefore advocate for the direct
measurement of chemical abundances with relatively few
free parameters, such as is described in this work, to be
used in conjunction with physical intuition derived from
the more complex thermal-chemical models.
It is interesting to note that the precision of derived
abundance profiles is primarily determined by the avail-
ability of high-quality spectra across a wide range of ex-
citation energies, as well as by the knowledge of the over-
all disk structure, especially the gas temperature. Thus,
more precise measurements can be obtained as spectro-
graphs with higher spectral resolution and sensitivity be-
come available, irrespective of the instrumental spatial
resolution (provided the disk structure is reasonably well
characterized) -- though of course instruments that pro-
vide detailed kinematic profiles of lines with a significant
range of excitation energies will provide the most strin-
gent probes. This technique thus provides an ideal means
to study chemical structure at the size scales relevant to
planet formation.
With a properly chosen spectral suite, we have demon-
strated that it is possible to probe molecular abundance
distributions on AU scales, and it is worth emphasiz-
ing that this method is highly complementary to the
capabilities of ALMA. At its longest baselines (16 km),
the full ALMA will be able to resolve a nearby disk
(140 pc) on AU scales at its shortest operational wave-
length, ∼400 µm, in dust emission. However, due to the
quantum-limited nature of heterodyne receivers and the
available collecting area, it will be a severe challenge even
for ALMA to robustly image the warm molecular gas in-
side of 10 AU -- the principle formation region of terres-
trial and giant planets according to core accretion theory.
Thus, for the foreseeable future, multi-wavelength meth-
ods (which can incorporate ALMA spectral data cubes)
offer perhaps the best means of deriving the molecular
abundance patterns in planet-forming environments. Ta-
11
Volatile line tracers at different disk radii
Table 3
Wavelength Radius
Facility
µm
2-5
6-10
10-35
40 - 200
200- 30000
AU
0.1
0.1- 1
0.5- 5
1-100
>50
NIRSPEC, CRIRES
SOFIA FORECAST
Spitzer IRS
Herschel PACS
HIFI, ALMA
Figure 9. The radial and vertical locations bounding water vapor
emission from the TW Hya disk. The boxes mark the 15% and 85%
cumulative radial line flux limits (vertical lines) and the heights
where 15% and 85% of the line flux arises from each vertical column
(horizontal lines). The gas density of the disk is shown in greyscale,
and the location of the Av ∼1 gas/dust decoupling transition (Tg =
Td) is shown by the dotted line. As a guide to the eye, the nH2 =
109 cm−3 density contours are labeled by the solid line. Excitation
and dust/line opacity are principally responsible for the lack of
sensitivity to the cold, dense disk midplane (c.f. Figure 2).
ble 3 presents the general selectivity of each wavelength
window for disks around Sun-like stars.
Although current observational S/N levels limit chemi-
cal abundance studies to a few species, this technique can
in principle be applied to the study of any molecule whose
transitions cover sufficient excitation space. The gas
phase distribution of condensible species provides partic-
ularly valuable information on the temperature, density,
chemical and transport patterns in disks; and used to-
gether, observations of a wide variety of molecules are
best suited to the isolation and thus understanding of
particular disk processes. To combine both infrared and
(sub)mm data on the inner/outer disk, polar species with
greatly differing sublimation temperatures are preferred.
The most obvious candidates, beyond water vapor and
OH, are CO and HCN, which are widely detected in sur-
veys of protoplanetary disks. The former provides a fidu-
cial chemical marker that only becomes significantly af-
fected by freeze-out at very low temperatures, while the
abundance drop for HCN across the snow line may well
be sensitive to the C/O ratio in the gas.
6. CONCLUSIONS
This paper uses multi-wavelength spectra to probe the
water vapor distribution in protoplanetary disks, and
demonstrates the first application of this method to in-
12
vestigate the water vapor abundance from 0.5-200 AU in
the transitional disk TW Hya. Our modeling shows that
there is a narrow region between 4-4.2 AU where water
vapor is warm (∼220 K) and optically thick, resulting in
a high abundance (XH2O∼10−4). Outside the snow-line,
the water vapor column density in the disk atmosphere
decreases dramatically over a scale length of less than
0.5 AU due to freeze out, resulting in a vertically inte-
grated vapor abundance of XH2O∼10−11 at large radii.
facilities
from the near-IR to radio frequencies can generate a rich
molecular data set of the emission lines from protoplan-
etary disks. We expect the multi-wavelength molecular
mapping method will soon be applied to a substantial
ensemble of disks and molecular species.
Both current and near-term astronomical
7. ACKNOWLEDGEMENTS
This work is based in part on observations made with
the Spitzer Space Telescope, which is operated by the
Jet Propulsion Laboratory, California Institute of Tech-
nology under a contract with NASA. Herschel is an ESA
space observatory with science instruments provided by
European-led Principal Investigator consortia and with
important participation from NASA. Funding for this
work (to G. Blake) was provided by the NASA Astrobiol-
ogy/Origins of Solar Systems programs and NSF Astron-
omy & Astrophysics. C. Salyk gratefully acknowledges
NOAO Leo Goldberg Postdoctoral Fellowship support.
We also thank the anonymous referee for a thorough and
insightful report, which helped to improve this paper.
REFERENCES
Aikawa, Y., Miyama, S. M., Nakano, T., & Umebayashi, T. 1996,
ApJ, 467, 684
Akeson, R. L., Millan-Gabet, R., Ciardi, D. R., Boden, A. F.,
Sargent, A. I., Monnier, J. D., McAlister, H., ten Brummelaar,
T., Sturmann, J., Sturmann, L., & Turner, N. 2011, ApJ, 728,
96
Andrews, S. M., Wilner, D. J., Espaillat, C., Hughes, A. M.,
Dullemond, C. P., McClure, M. K., Qi, C., & Brown, J. M.
2011, ApJ, 732, 42
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., Rosenfeld,
K. A., Oberg, K. I., Birnstiel, T., Espaillat, C., Cieza, L. A.,
Williams, J. P., Lin, S.-Y., & Ho, P. T. P. 2012, ApJ, 744, 162
Armitage, P. J. 2011, ARA&A, 49, 195
Arnold, T. J., Eisner, J. A., Monnier, J. D., & Tuthill, P. 2012,
ApJ, 750, 119
Banzatti, A., Meyer, M. R., Bruderer, S., Geers, V., Pascucci, I.,
Lahuis, F., Juh´asz, A., Henning, T., & ´Abrah´am, P. 2012, ApJ,
745, 90
Baulch, D. L., ed. 1972, Evaluated kinetic data for high
temperature reactions (CRC Press)
Bergin, E. A., Cleeves, L. I., Gorti, U., Zhang, K., Blake, G. A.,
Green, J., Andrews, S. M., Evans, II, N. J., Henning, T.,
Oberg, K., Pontoppidan, K. M., Salyk, C., & van Dishoeck,
E. F. 2013, Nature, 393, 7433
Bergin, E. A., Hogerheijde, M. R., Brinch, C., & Fogel, J. e. a.
2010, A&A, 521, L33+
Bethell, T., & Bergin, E. 2009, Science, 326, 1675
Birnstiel, T., Dullemond, C. P., & Brauer, F. 2009, Astronomy
and Astrophysics, 503, L5
Blum, J., & Wurm, G. 2008, ARA&A, 46, 21
Calvet, N., D'Alessio, P., Hartmann, L., Wilner, D., Walsh, A., &
Sitko, M. 2002, ApJ, 568, 1008
Carr, J. S., & Najita, J. R. 2008, Science, 319, 1504
Ciesla, F. J., & Cuzzi, J. N. 2006, Icarus, 181, 178
D'Alessio, P., Calvet, N., & Hartmann, L. 2001, ApJ, 553, 321
Dodson-Robinson, S. E., Willacy, K., Bodenheimer, P., Turner,
N. J., & Beichman, C. A. 2009, Icarus, 200, 672
Dullemond, C. P., & Dominik, C. 2004, A&A, 417, 159
Eisner, J. A., Chiang, E. I., & Hillenbrand, L. A. 2006, ApJ, 637,
L133
Fedele, D., Bruderer, S., van Dishoeck, E. F. v., Herczeg, G. J.,
Evans, N. J., Bouwman, J., Henning, T., & Green, J. 2012,
A&A, 544, L9
Frisch, P. C., & Slavin, J. D. 2003, ApJ, 594, 844
Garaud, P., & Lin, D. N. C. 2007, ApJ, 654, 606
Glassgold, A. E., Meijerink, R., & Najita, J. R. 2009, ApJ, 701,
142
Glassgold, A. E., Najita, J. R., & Igea, J. 2004, ApJ, 615, 972
Gomes, R., Levison, H. F., Tsiganis, K., & Morbidelli, A. 2005,
Nature, 435, 466
Gorti, U., Hollenbach, D., Najita, J. R., & Pascucci, I. 2011, ApJ,
735, 90
Hayashi, C. 1981, Progress of Theoretical Physics Supplement,
70, 35
Heinzeller, D., Nomura, H., Walsh, C., & Millar, T. J. 2011, ApJ,
731, 115
Hogerheijde, M. R., Bergin, E. A., Brinch, C., Cleeves, L. I.,
Fogel, J. K. J., Blake, G. A., Dominik, C., Lis, D. C., Melnick,
G., Neufeld, D., Panic, O., Pearson, J. C., Kristensen, L., Yldz,
U. A., & van Dishoeck, E. F. v. 2011, Science, 334, 338
Hughes, A. M., Wilner, D. J., Calvet, N., D'Alessio, P., Claussen,
M. J., & Hogerheijde, M. R. 2007, ApJ, 664, 536
Johansen, A., Oishi, J. S., Mac Low, M.-M., Klahr, H., Henning,
T., & Youdin, A. 2007, Nature, 448, 1022
Johansen, A., Youdin, A., & Mac Low, M.-M. 2009, ApJ, 704, L75
Kamp, I., & Dullemond, C. P. 2004, ApJ, 615, 991
Kennedy, G. M., & Kenyon, S. J. 2008, ApJ, 673, 502
Kretke, K. A., & Lin, D. N. C. 2007, ApJ, 664, L55
Lecar, M., Podolak, M., Sasselov, D., & Chiang, E. 2006, ApJ,
640, 1115
Low, F. J., Smith, P. S., Werner, M., Chen, C., Krause, V., Jura,
M., & Hines, D. C. 2005, ApJ, 631, 1170
Mamajek, E. E. 2005, ApJ, 634, 1385
Meeus, G., Montesinos, B., Mendigut´ıa, I., Kamp, I., Thi, W.-F.,
& team, t. G. H. 2012, arXiv.org, astro-ph.GA
Meijerink, R., Pontoppidan, K. M., Blake, G. A., Poelman, D. R.,
& Dullemond, C. P. 2009, ApJ, 704, 1471
Najita, J. R., ´Ad´amkovics, M., & Glassgold, A. E. 2011, ApJ,
743, 147
Najita, J. R., Carr, J. S., Strom, S. E., Watson, D. M., Pascucci,
I., Hollenbach, D., Gorti, U., & Keller, L. 2010, ApJ, 712, 274
Pollack, J. B., Hollenbach, D., Beckwith, S., Simonelli, D. P.,
Roush, T., & Fong, W. 1994, ApJ, 421, 615
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J.,
Podolak, M., & Greenzweig, Y. 1996, Icarus, 124, 62
Pontoppidan, K. M., Blake, G. A., van Dishoeck, E. F., Smette,
A., Ireland, M. J., & Brown, J. 2008, ApJ, 684, 1323
Pontoppidan, K. M., Meijerink, R., Dullemond, C. P., & Blake,
G. A. 2009, ApJ, 704, 1482
Pontoppidan, K. M., Salyk, C., Blake, G. A., & Kaufl, H. U.
2010a, ApJ, 722, L173
Pontoppidan, K. M., Salyk, C., Blake, G. A., Meijerink, R., Carr,
J. S., & Najita, J. 2010b, ApJ, 720, 887
Qi, C., Ho, P. T. P., Wilner, D. J., Takakuwa, S., Hirano, N.,
Ohashi, N., Bourke, T. L., Zhang, Q., Blake, G. A.,
Hogerheijde, M. R., Saito, M., Choi, M., & Yang, J. 2004, ApJ,
616, L11
Raymond, S. N., Quinn, T., & Lunine, J. I. 2004, Icarus, 168, 1
Rettig, T. W., Haywood, J., Simon, T., Brittain, S. D., & Gibb,
E. 2004, ApJ, 616, L163
Riviere-Marichalar, P., M´enard, F., Thi, W.-F., Kamp, I.,
Montesinos, B., Meeus, G., Woitke, P., Howard, C., Sandell, G.,
Podio, L., Dent, W. R. F., Mendigut´ıa, I., Pinte, C., White,
G. J., & Barrado, D. 2012, Astronomy and Astrophysics, 538,
L3
Rosenfeld, K. A., Qi, C., Andrews, S. M., Wilner, D. J., Corder,
S. A., Dullemond, C. P., Lin, S.-Y., Hughes, A. M., D'Alessio,
P., & Ho, P. T. P. 2012, ApJ, 757, 129
Rucinski, S. M., & Krautter, J. 1983, A&A, 121, 217
Salyk, C., Blake, G. A., Boogert, A. C. A., & Brown, J. M. 2007,
ApJ, 655, L105
-- . 2009, ApJ, 699, 330
Salyk, C., Pontoppidan, K. M., Blake, G. A., Lahuis, F., van
Dishoeck, E. F., & Evans, II, N. J. 2008, ApJ, 676, L49
Salyk, C., Pontoppidan, K. M., Blake, G. A., Najita, J. R., &
Carr, J. S. 2011, ApJ, 731, 130
Stevenson, D. J., & Lunine, J. I. 1988, Icarus, 75, 146
Supulver, K. D., & Lin, D. N. C. 2000, Icarus, 146, 525
Thi, W.-F., Mathews, G., M´enard, F., Woitke, P., Meeus, G.,
Riviere-Marichalar, P., Pinte, C., Howard, C. D., Roberge, A.,
Sandell, G., Pascucci, I., Riaz, B., Grady, C. A., Dent,
W. R. F., Kamp, I., Duchene, G., Augereau, J.-C., Pantin, E.,
Vandenbussche, B., Tilling, I., Williams, J. P., Eiroa, C.,
Barrado, D., Alacid, J. M., Andrews, S., Ardila, D. R., Aresu,
G., Brittain, S., Ciardi, D. R., Danchi, W., Fedele, D., de
Gregorio-Monsalvo, I., Heras, A., Huelamo, N., Krivov, A.,
Lebreton, J., Liseau, R., Martin-Zaidi, C., Mendigut´ıa, I.,
Montesinos, B., Mora, A., Morales-Calderon, M., Nomura, H.,
Phillips, N., Podio, L., Poelman, D. R., Ramsay, S., Rice, K.,
Solano, E., Walker, H., White, G. J., & Wright, G. 2010, A&A,
518, L125
13
Tielens, A. G. G. M., & Hagen, W. 1982, A&A, 114, 245
Uchida, K. I., Calvet, N., Hartmann, L., Kemper, F., Forrest,
W. J., Watson, D. M., D'Alessio, P., Chen, C. H., Furlan, E.,
Sargent, B., Brandl, B. R., Herter, T. L., Morris, P., Myers,
P. C., Najita, J. R., Sloan, G. C., Barry, D. J., Green, J.,
Keller, L. D., & Hall, P. 2004, ApJS, 154, 439
Walsh, C., Nomura, H., Millar, T. J., & Aikawa, Y. 2012, ApJ,
747, 114
Webb, R. A., Zuckerman, B., Platais, I., Patience, J., White,
R. J., Schwartz, M. J., & McCarthy, C. 1999, ApJ, 512, L63
Weidenschilling, S. J. 1997, Icarus, 127, 290
Weintraub, D. A., Sandell, G., & Duncan, W. D. 1989, ApJ, 340,
L69
Woitke, P., Kamp, I., & Thi, W.-F. 2009, A&A, 501, 383
|
1704.01121 | 2 | 1704 | 2017-11-12T00:18:07 | A companion on the planet/brown dwarf mass boundary on a wide orbit discovered by gravitational microlensing | [
"astro-ph.EP",
"astro-ph.SR"
] | We present the discovery of a substellar companion to the primary host lens in the microlensing event MOA-2012-BLG-006. The companion-to-host mass ratio is 0.016, corresponding to a companion mass of $\approx8~M_{\rm Jup} (M_*/0.5M_\odot)$. Thus, the companion is either a high-mass giant planet or a low-mass brown dwarf, depending on the mass of the primary $M_*$. The companion signal was separated from the peak of the primary event by a time that was as much as four times longer than the event timescale. We therefore infer a relatively large projected separation of the companion from its host of $\approx10~{\rm a.u.}(M_*/0.5M_\odot)^{1/2}$ for a wide range (3-7 kpc) of host star distances from the Earth. We also challenge a previous claim of a planetary companion to the lens star in microlensing event OGLE-2002-BLG-045. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. paper
November 15, 2017
c(cid:13)ESO 2017
A companion on the planet/brown dwarf mass
boundary on a wide orbit discovered by
gravitational microlensing
R. Poleski1, 2, A. Udalski2, I. A. Bond3, J. P. Beaulieu4, 5, 6, C. Clanton7, S. Gaudi1,
M. K. Szyma´nski2, I. Soszy´nski2, P. Pietrukowicz2, Szymon Kozłowski2, J. Skowron2,
and,
Ł. Wyrzykowski2, K. Ulaczyk8, 2,
and,
D. P. Bennett9, T. Sumi10, D. Suzuki9, N. J. Rattenbury11, N. Koshimoto10, F. Abe12,
Y. Asakura12, R. K. Barry9, A. Bhattacharya9, 13, M. Donachie11, P. Evans11, A. Fukui14,
Y. Hirao10, Y. Itow12, M. C. A. Li11, C. H. Ling3, K. Masuda12, Y. Matsubara12,
Y. Muraki12, M. Nagakane10, K. Ohnishi15, C. Ranc9, To. Saito16, A. Sharan11,
D. J. Sullivan17, P. J. Tristram18, T. Yamada10, T. Yamada19, A. Yonehara19,
and,
V. Batista4, J. B. Marquette4
(Affiliations can be found after the references)
Received April 3, 2017; accepted May 4, 2017
ABSTRACT
We present the discovery of a substellar companion to the primary host lens in the microlensing event
MOA-2012-BLG-006. The companion-to-host mass ratio is 0.016, corresponding to a companion mass of
≈ 8 MJup(M∗/0.5M(cid:12)). Thus, the companion is either a high-mass giant planet or a low-mass brown dwarf,
depending on the mass of the primary M∗. The companion signal was separated from the peak of the primary
event by a time that was as much as four times longer than the event timescale. We therefore infer a relatively
large projected separation of the companion from its host of ≈ 10 a.u.(M∗/0.5M(cid:12))1/2 for a wide range (3-7
kpc) of host star distances from the Earth. We also challenge a previous claim of a planetary companion to the
lens star in microlensing event OGLE-2002-BLG-045.
Key words. gravitational lensing: micro – planetary systems – brown dwarfs – instrumentation: high angular
resolution
Article number, page 1 of 16
7
1
0
2
v
o
N
2
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
2
1
1
0
.
4
0
7
1
:
v
i
X
r
a
Poleski et al.: Planet/brown dwarf on a wide orbit
1. Introduction
Brown dwarfs and planets are intrinsically faint objects and different detection techniques have to
be used to explore a wide range of properties of these sub-stellar objects. Every detection technique
has its own limitations and leads to a different kind of information when a new object is detected.
Despite the large number of observational and theoretical studies (e.g., Beichman et al. 2014; Chau-
vin et al. 2015; Foreman-Mackey et al. 2016; Wilson et al. 2016), we are still far from a detailed
understanding of the demographics of brown dwarf and planet populations that are also compan-
ions to stars. There is even a lack of consensus on the appropriate border line between planets and
brown dwarfs (Boss et al. 2003; Grether & Lineweaver 2006; Spiegel et al. 2011; Chabrier et al.
2014). The obvious way to increase our knowledge of sub-stellar mass objects is by discovering
more objects and, in particular, by discovering and characterizing objects that question our current
understanding of planet and brown dwarf formation and evolution.
Here, we present the discovery of a binary system MOA-2012-BLG-006L with a mass ratio of
0.016 and projected separation of roughly 10 a.u. Both components of the system were detected
using the gravitational microlensing method. The advantage of this method is that it is sensitive to
the mass of the objects, rather than their luminosity. As a result, microlensing enables the discovery
of systems that are inaccessible to other techniques. First, the system distance of a few kpc pre-
vents the detection of light from the lower-mass component via direct imaging. Second, the radial
velocity signal of a long-period, low-mass companion to the faint host is out of reach of current
techniques. Finally, the projected separation of about 10 a.u. results in the extremely low probabil-
ity of observing the transit, even if a population of similar systems was observed using photometric
methods. We note that there are three other systems that were discovered using microlensing and
contain either brown dwarfs or planets: MOA-2007-BLG-197L (Ranc et al. 2015), MOA-2010-
BLG-073L (Street et al. 2013), and MOA-2011-BLG-322L (Shvartzvald et al. 2014). These three
systems have smaller separations and higher mass ratios compared to the system reported here,
MOA-2012-BLG-006L. The distribution of mass ratios for binary lens microlensing events was re-
cently investigated by Shvartzvald et al. (2016). They found that the mass ratio distribution shows
the minimum and this minimum is close to the mass ratio of MOA-2012-BLG-006L (0.016).
Explaining the formation of the MOA-2012-BLG-006L system poses significant challenges.
The mass of the protoplanetary disc is typically 0.002-0.006 of the host mass (Andrews et al.
2013), hence, any planet that forms in a protoplanetary disc that follows this observational trend
cannot have a larger mass ratio. If the protoplanetary disc in MOA-2012-BLG-006L had a mass
ratio close to the typical values, then the lower-mass object should be classified as a brown dwarf.
However, there is a wide range of measured disc masses at fixed host mass (Andrews et al. 2013).
In the extreme cases, estimated disc masses are close to 0.2 of the host mass (Andrews et al. 2009).
In these extreme cases, the total mass of the disc is sufficient to form planets with mass ratios
similar to MOA-2012-BLG-006Lb. The planetary formation scenario poses an additional question:
how did a planet so massive end up on an orbit that is at least eight times larger than the snow line
Article number, page 2 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
distance (≈ 1.3 a.u. in this case)? Most massive planets formed by core accretion should do so just
beyond the snow line, where the protoplanetary disc is still relatively dense and ices can condense.
Furthermore, if the planet formed via gravitational instability, we might expect it to be on an orbit
wider by a factor of a few in semimajor axis (which, depending on the projection, it may actually
be; Dodson-Robinson et al. 2009)
In the following Section we describe photometric observations leading to the discovery of
MOA-2012-BLG-006Lb. In Section 3 we analyze photometric data and derive the system prop-
erties using a Galactic model. The degeneracies in the microlensing model fitting are described in
detail. The following section presents high-resolution observations of the event. Section 5 discusses
another microlensing event (OGLE-2002-BLG-045) that showed a possible anomaly that could be
fitted with a planetary model. We conclude that the anomaly was not real and there is no evidence
for a planet. We end with conclusions.
2. Photometric observations
The microlensing event MOA-2012-BLG-006 was announced by the MOA group (Microlensing
Observations in Astrophysics; Bond et al. 2001) on Feb 9, 2012 (HJD(cid:48) ≡ HJD−2450000 = 5967.3)
at (R.A., Dec.) = (18h01m46s.31, −29◦06(cid:48)31(cid:48)(cid:48).6) (Galactic coordinates l ≈ 1◦.64, b ≈ −3◦.13). The
event was found very early during the bulge observing season. During that time, bulge is visible
only for a short time each night from any single site. The chances of discovering planets so early
during the bulge observing season are low and most of the follow-up surveys do not start their
normal operations before about a month later. Hence, survey observations are the only way to find
planets that show their signatures so early in the season. The same event was alerted by the OGLE
survey (Optical Gravitational Lensing Experiment; Udalski 2003) on Feb 13, 2012 (HJD(cid:48) = 5971)
in the first batch of the microlensing events in 2012 and labeled OGLE-2012-BLG-0022. The
OGLE and the MOA survey telescopes are well separated in geographic longitude, which allows
coverage of different parts of the light curve. Below we describe the datasets produced by the two
surveys.
The main photometric dataset comes from the OGLE survey, which uses a 1.3m telescope
located at Las Campanas Observatory (Chile). The telescope is equipped with the 32-CCD mosaic
camera that gives a 1.4 deg2 field of view (Udalski et al. 2015). There are eight photometric epochs
during the anomaly (i.e., HJD(cid:48) between 5960 and 5971) and 2306 more measurements during the
2012 bulge season – see light curve in Figure 1. We also included 813 datapoints from 2011 in order
to ensure that the baseline brightness is correctly measured. The OGLE survey performs most of
the observations in the I-band and we use only these data for fitting. The V-band data do not cover
the anomalous part of the light curve and are only used to derive source properties. The photometry
was performed using the Difference Image Analysis (DIA) method (Alard 2000; Wo´zniak 2000).
The photometric uncertainties were corrected using the prescription presented by Skowron et al.
(2016). There are two relatively bright field stars that are very close to the event: 1.1 and 1.4 arcsec
Article number, page 3 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Fig. 1. Light curve of MOA-2012-BLG-006 = OGLE-2012-BLG-0022. OGLE and MOA data are marked
using black and red points, respectively. Panel a presents the 2012 light curve with both subevents. Panel b
shows the model residuals. Panel c gives zoom-in on the anomaly. Panel d presents the source trajectory (black
line) relative to the planetary caustic (blue curve), which is at the origin of the coordinate system. The host
star is located at (θx/θE, θy/θE) = (4.17, 0.0). Source positions from one OGLE night and two MOA nights are
marked and are aligned with photometry shown in panel c. The circles have a radius of ρ.
away with an I-band brightness of 17.6 and 16.2 mag, respectively. The two stars can affect the
photometry of the event. Indeed, we found that seeing variations marginally influence brightness
measurements – the target gets fainter by 0.005 mag for an increase in seeing FWHM of 1 arcsec.
This effect was subtracted from the OGLE data.
The second dataset used comes from the MOA survey. The MOA survey operates a 1.8m tele-
scope situated at Mt. John Observatory (New Zealand). The filter used for observations is a custom
wide-band optical filter. The camera consists of ten CCD detectors and gives a 2.2 deg2 field of
view (Sako et al. 2008). The MOA observing site has poorer weather and seeing conditions as
compared to the OGLE site, but enables observations of microlensing events when Galactic bulge
is invisible from Chile. Photometry was performed using the DIA method (Bond et al. 2001). MOA
data for the analyzed event are more affected by the variable seeing and additionally show depen-
dence of measured brightness on the airmass. Unfortunately, the specific way in which data are
Article number, page 4 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
affected changes over time. We see these changes both during the event, and during the other ob-
serving seasons. In order not to include the affected data in the fit, we restricted the MOA dataset
to the fourteen epochs that are closest to the anomaly, i.e., from HJD(cid:48) = 5961.1 to 5969.2. Sim-
ilar issues with a nearby star contaminating photometry of the microlensing event were faced by
Gould et al. (2014) who analyzed the event OGLE-2013-BLG-0341. We note that the remaining
MOA data will not improve accuracy of the fitted event properties, because with the exception of
the anomaly, the event was of a low magnification and during bulge observing season the OGLE
cadence of 20 min is more than sufficient to characterize the light curve. We also checked that the
other bulge photometric survey operating at that time – VISTA Variables in the Vía Láctea (VVV;
Minniti et al. 2010) – did not collect any data of this field during the anomaly.
3. Analysis
The light curve of MOA-2012-BLG-006 resembles a superposition of two point-source/point-lens
microlensing events. Light curves of this type can be produced in two physically different scenarios
(Gaudi 1998). First, the lens can be a single object and the source can be a binary system leading
to two subevents with the same Einstein timescale tE. We tried to fit the binary source model to the
observed light curve, but could not find a good fit. After rejecting the binary source model, we are
left with only one other possibility – the lens is a binary system and a single source is magnified
(Gaudi 1998). If the two subevents are not significantly affected by the caustics (curves on which
point source magnification is infinite), then the mass ratio of the two lens components is a square
of the tE ratio of the subevents. In the present case, a simple examination of the light curve by eye
suggests that the lower mass object is either a planet or a brown dwarf if the host is a typical main
sequence star. The first subevent has a higher magnification, even though it was caused by the lower
mass object. The magnification of the subevent depends primarily on the impact parameter, not the
lens mass.
To fit the microlensing model we evaluated magnification using the inverse ray shooting method
for the highest magnified points and the hexadecapole approximation (Gould 2008; Pejcha & Hey-
rovský 2009) for the adjacent parts of the light curve. We used the complex polynomial root solver
by Skowron & Gould (2012). Based on Claret & Bloemen (2011) and source properties derived
from the initially fitted model, we set the limb darkening coefficients to ΓI = 0.502 (uI = 0.602)
and ΓMOA = 0.588 (uMOA = 0.681). We note that initial fitting was performed to the OGLE data
only, but the results do not qualitatively differ from fits to the OGLE and MOA data.
We first tried to fit the model using Monte Carlo Markov Chain (MCMC) that is typically used
for the analysis of the microlensing events. We used MCMC implementation by Dong et al. (2007)
and Poleski et al. (2014). Even though we run the MCMC multiple times with a number of settings,
MCMC failed to produce a converging chain and hence we could not use it to fit a microlensing
model. The triangle (or corner) plot showed that almost all two-parameter marginalized χ2 hy-
persurfaces had approximately ellipsoidal shapes but still the chain was not converging. The only
Article number, page 5 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
exceptions were χ2 hypersurfaces where one of the parameters was the angular source radius rela-
tive to the Einstein ring radius (ρ). We fully understood the reason for failure in the MCMC runs
only after transforming the microlensing model was transformed to a different set of parameters.
A default set of seven parameters that describes a binary lens model consists of: the three point
lens parameters, ρ, and the three parameters that describe the lens companion. The three point lens
parameters are: t0 – the epoch of minimum source-lens separation, u0 – the minimum separation
relative to Einstein ring radius θE, and tE. The lens companion is described by: α – the angle be-
tween the lens axis and the source trajectory, s – the separation of the lens components relative to
θE, and q – the mass ratio. In addition to the binary lens parameters that completely describe the
magnification, the model also contains source and blending fluxes for each photometric system.
The default binary lens parametrization is not optimal for fitting all the microlensing events be-
cause the parameters α, q, and in many cases s are not directly constrained by the light curves, in
the sense that the observable properties of the microlensing event are not directly relatable to these
parameters (Cassan 2008; Sumi et al. 2010; Skowron et al. 2011; Kains et al. 2012).
The event MOA-2012-BLG-006 shows two well-separated subevents and their parameters
(maximum magnification, its epoch, and the length of the subevent) are well-constrained by the
data, assuming the blending flux is known. Hence for fitting, we used, instead of the default set of
parameters, ρ and the three point lens parameters measured separately for each of the components
(tE) or relative to their caustics (t0 and u0). We note that in this two-component parameterization
either u0,1 or u0,2 has to be a signed quantity in order to make a distinction between the source
passing both caustics on same or opposite sides (unlike in a point-source/point-lens model without
parallax). The conversion between both binary lens parameterizations is based on simple geometry
and the equation for distance between the central and the planetary caustics: s − s−1 (Han 2006).
In the two-component parametrization, we can easily find and understand the very significant
model degeneracies. We present the slice of χ2 hypersurface in Figure 2. There are three local
minima for different source trajectories and ρ values. The best-fitting model has a ρ of ≈ 0.012 and
the source trajectory that passes each caustic on a different side (see Figure 1). The second best-
fitting model has a much smaller source (ρ ≈ 0.001 or even smaller) and the source passing through
the center of the planetary caustic. The small source causes the characteristic U-shaped light curve
(see the panel d of Figure 2) but both high-magnification parts of the light curve (when the source
crosses the caustic) that reach I-band magnitude of 12.3 are predicted to have occurred during
the time when no data were taken. In this model, the brightest OGLE data point is taken close
to the middle of the U-shaped trough. A Bayesian argument suggests that a model that predicts
a large brightness variation during a time when no data were taken is a priori unlikely, although
this argument alone cannot rule out such a model. However, we can exclude this model because it
predicts unreasonably large relative lens-source geocentric proper motion µ = θE/tE = θ(cid:63)/(ρtE) on
the order of 100 mas yr−1, where θ(cid:63) is the angular radius of the source star equal to 5.68± 0.34 µas
(see Section 3.1). In the third solution, the source passes both caustics on the same side and has
Article number, page 6 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Fig. 2. Degenerate solutions for MOA-2012-BLG-006. Top center panel d presents anomaly part of light
curves for three modes A, B, and C. All three light curves predict the second subevent at HJD(cid:48) = 6047.
Remaining large panels a, b, and c show the projection of the marginalized χ2 hypersurface. Red, orange,
yellow, green, and blue points correspond to ∆χ2 of < 1, < 4, < 9, < 16, and < 25, respectively. Letters A,
B, and C mark the three modes. We note that u0,1 > 0 means that the companion and the host caustics are
passed on opposite sides. The data presented on this plot are not the result of a single run, but a compilation
of many simulations and were obtained using OGLE data only. Similar plot for OGLE and MOA data does
not show significant differences except different levels of ∆χ2. In particular, MOA data at HJD(cid:48) = 5965.2 and
5967.2 significantly contribute to preference of mode A over mode C. The small panels on the right e, f, and
g show source trajectories of the three solutions relative to planetary caustic (blue curve). Crosses mark the
source positions at epochs when OGLE data were taken. Circles have a radius of ρ. In the case of solution B,
the source size of ρ = 0.00016 is too small to be seen.
ρ ≈ 0.006. We reject this solution because it is worse than the first solution by ∆χ2 = 5.8 (∆χ2 =
51.3 if MOA data are included).
Ultimately, we decided not to use the usual MCMC algorithm for fitting the model, but instead
apply an alternative algorithm that is more suited to explore degenerate multidimensional distribu-
tions – Multimodal Ellipsoidal Nested Sampling or MultiNest. The algorithm is described in detail
and implemented by Feroz & Hobson (2008) and Feroz et al. (2009, 2013). In brief, MultiNest
approximates a volume of parameter space for which χ2 is below some limiting value χ2
0 by a set of
N points. We used N = 5000 here for the final fitting, but reasonably good exploration of parameter
space is achieved even for N = 500 when considering only OGLE data. Both parameter estimation
and model selection result from a single simulation in which χ2
0 is reduced from one step to the
next. In every step, one of the N points with the highest χ2 is replaced by a point that has lower
χ2 and was found by trial-and-error. The trial-and-error procedure randomly samples a union of
ellipsoids enclosing the N points according to the prior (which was uniform in linear parameters).
For the final model fitting, we used both OGLE and MOA data and assumed ρ > 0.007. The ρ
constraint is equivalent to assuming that µ < 15 mas yr−1. The Galactic simulation described below
Article number, page 7 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Table 1. Double lens model parameters
d
Quantity Unit
t0
u0
tE
ρ
α
s
q
Fs/Fbase
χ2/dof
deg
Value
6046.87 ± 0.12
1.432 ± 0.032
20.69 ± 0.35
0.0119+0.0016
−0.0023
20.17 ± 0.20
4.405 ± 0.069
0.01650 ± 0.00055
0.981 ± 0.054
2898.21/3130
Notes. The parameter Fs/Fbase indicates ratio of source flux to baseline flux in the I-band. The value of u0 is
greater than one, hence, the host subevent would not normally be counted for the optical depth calculations.
(with only tE constrained) gives a probability of µ > 15 mas yr−1 to be < 0.001. We present the final
model parameters obtained using MultiNest in Table 1. Even though the fitting was performed in
the two-component parameterization, all the parameters in default parametrization except ρ show
symmetric posterior distributions. We note that the event MOA-2012-BLG-006 was used in a sta-
tistical analysis of the exoplanet mass ratio function by Suzuki et al. (2016). The parameters used
there (tE = 21.13 d, u0 = 1.3, q = 0.01614, s = 4.32) slightly differ from those found in the present
work. The Suzuki et al. (2016) analysis was performed independently from the present analysis.
The problems with fitting the microlensing models described above are primarily caused by the
poor coverage of the anomalous part of the light curve. Similar problems frequently appear during
the analysis of poorly sampled anomalies. Jaroszynski & Paczynski (2002) claimed that a single
point anomaly in OGLE-2002-BLG-055 could be explained by the planetary model. Later, Gaudi
& Han (2004) showed that there is a plethora of models with non-planetary mass ratios that can
fit the same light curve. Analysis of the planetary event MOA-2007-BLG-192 revealed degenerate
cusp approach and caustic crossing solutions that could not be efficiently sampled by MCMC runs
(Bennett et al. 2008) because of the huge number of steps needed to cross the χ2 barrier between
them. We predict that MultiNest can solve the remaining problems in analysis of these, and other,
poorly sampled events.
Direct measurement of the lens mass, distance, and projected separation of the lens components
requires microlensing parallax πE (Gould 2000) to be measured. We cannot measure or even put
meaningful constraints on πE for MOA-2012-BLG-006 because the host subevent is too short and
the value of u0 is too large. The companion subevent could reveal the parallax signal only if it
was sampled at a much higher cadence. Without a parallax measurement, we have to use source
properties and Bayesian priors using a Galactic model to constrain the lens mass, distance, and
projected separation of components.
3.1. Source properties
The lens mass and distance are crucial parameters for determining the nature of the lens system.
These parameters cannot be directly derived from only the microlensing parameters like tE and ρ.
Article number, page 8 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Microlensing events with finite source effects (and thus measured ρ) can only be used to estimate
the physical properties of the lens if we can measure the angular Einstein ring radius θE = θ(cid:63)/ρ
and also measure the microlensing parallax or the lens flux. Here the estimate of θ(cid:63) follows the
method by Yoo et al. (2004). First, we construct the color-magnitude diagram for stars lying close
to the event as presented in Figure 3. Second, we measure the properties of the red clump (RC):
IRC = 15.767±0.017 mag and (V−I)RC = 2.024±0.007 mag. Third, by comparing these values with
theoretical values, IRC,0 = 14.381 mag (found by interpolation of Table 2 from Nataf et al. 2013)
and (V − I)RC,0 = 1.06 mag (Bensby et al. 2011), we find extinction AI = 1.386 mag and reddening
E(V − I) = 0.964 mag. Fourth, we correct the source's unmagnified brightness (Is = 16.247 mag
and Vs = 18.390 mag) for extinction and obtain: Is,0 = 14.861 mag and Vs,0 = 16.040 mag. Fifth,
the extinction-corrected brightness in visual bands is transformed to the near-infrared brightness
of Ks,0 = 13.314 mag based on the intrinsic colors of giant stars (Bessell & Brett 1988). Sixth,
the angular source radius of θ(cid:63) = 5.68 ± 0.36 µas is calculated using the relation between surface
brightness and (V − K) color by Kervella et al. (2004). This procedure results in θE = θ(cid:63)/ρ =
0.489+0.126
−0.038 mas.
The event was also observed in J-band using adaptive optics (AO) at Keck telescope. Bessell
& Brett (1988) relations predict an extinction-free source brightness of Js,0 = 14.077 ± 0.085 mag.
The J-band extinction toward the event is AJ = 0.58± 0.16 mag (Gonzalez et al. 2012). Hence, we
predict observed source brightness of Js = 14.66 ± 0.19 mag.
3.2. Galactic model
To derive the physical properties of the lens, we simulated microlensing events using the modified
version of Galactic model by Clanton & Gaudi (2014) and we refer the reader to that paper for a
detailed description. The model includes lenses from Galactic disc with a double-exponential den-
sity profile and boxy Gaussian bulge. The line-of-sight projected velocity of the Earth is calculated
for a peak of the anomaly. The lens mass distribution is the same as in Sumi et al. (2011) model 1
limited to the main sequence lenses: power laws with α = 1.3 for 0.08 ≤ M/M(cid:12) < 0.7 and α = 2.0
for 0.7 ≤ M/M(cid:12) < 1.0. Sources are placed at a distance of 7.8 kpc. The ensemble of simulated
events is additionally weighted according to the measured tE and θE. No constraint on the lens flux
was applied. The resulting distributions are used to estimate the physical properties of the lens. We
used the Astropy package (Astropy Collaboration et al. 2013) to analyze the simulation.
Figure 4 and Table 2 present the posterior distributions of the event parameters as derived from
the Galactic model: Mh and Mc – mass of the host and companion, respectively, a⊥ – projected
separation of host and companion, Dl – distance to the lens, and µ – relative proper motion of
lens and source. We note that a⊥ = 10.2 a.u. corresponds to deprojected semi-major axis (for a
√
3/2a⊥ = 12.5 a.u. The
circular orbit with a random value of the cosine of the inclination) of a =
probability that Mc is above the frequently assumed minimum brown dwarf mass of 13MJ is 0.18.
Article number, page 9 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Fig. 3. Color-magnitude diagram for stars within 2(cid:48) around MOA-2012-BLG-006. Red circle marks centroid
of the red clump. The blue cross marks the position of the source.
Figure 4 also includes predictions of lens near-infrared brightness based on Dotter et al. (2008) 6
Gyr isochrone for [Fe/H] = 0.0 and Y = 0.27.
4. High-resolution observations
On July 18, 2013 (1.2 yr after peak of the event) we observed the microlensing event MOA-2012-
BLG-006 with Near Infrared Camera 2 (NIRC-2) AO system mounted on the Keck-II telescope.
We used the wide field (40(cid:48)(cid:48) × 40(cid:48)(cid:48)) camera with a pixel scale of 0.04 arcsec and J-band filter. We
took four frames with an exposure time of 3 × 10 seconds at each of the five dithered positions.
Article number, page 10 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Fig. 4. Posterior distributions of parameters derived from the Galactic model. The three lower panels show
predicted brightness of the lens in J, H, and Ks filters (2MASS system). The extinction of AJ = 0.58, AH =
0.33, and AKs
= 0.20 (Gonzalez et al. 2012) were assumed independent of the lens distance.
We corrected for dark and flat fields using standard procedure and stacked the images using SWarp
program from the AstrOmatic suite of astronomy tools (Bertin 2010). The full width at half max-
imum was 0.2 arcsec. The aperture photometry was performed by running SExtractor (Bertin &
Article number, page 11 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Table 2. Posterior statistics for event parameters using a Bayesian prior and a Galactic model. Mean values
and 1 − σ uncertainties are given.
Quantity Unit
M(cid:12)
Mh
Mc
MJ
Dl
kpc
a⊥
a.u.
mas/yr
µ
Value
0.49+0.27−0.23
8.4+4.6−3.9
5.3+0.8−1.3
10.2+1.8−2.4
7.69+1.1−0.76
Fig. 5. Keck AO image of the event. The cross marks the expected position.
Arnouts 1996) software. The photometric and astrometric calibration of the Keck images requires
additional data and for this purpose we used VVV data. The VVV survey observed in J, H, and K
bands at the 4m VISTA telescope at Paranal Observatory (Chile). To process VVV images we fol-
lowed the procedure described by Beaulieu et al. (2016) which includes calibration of photometry
and astrometry to 2MASS system (Skrutskie et al. 2006).
The source star is clearly identified on Keck image at the expected position (see Figure 5).
We note that there is no significant blend at the sub-arcsec separation. The VVV brightness is
JVVV = 14.72± 0.02, HVVV = 13.88± 0.02, and KVVV = 13.62± 0.03. Based on cross-identification
of the same stars in Keck and VVV data we estimate the error in absolute calibration of Keck
photometry of 0.015 mag. The brightness measured on Keck image and calibrated to VVV data is
JKeck = 14.70 ± 0.03.
We can compare the total object brightness measured from the Keck image with the lens and
source brightness estimated before. The fiducial lens mass from our Galactic model (0.49 M(cid:12))
corresponds to absolute brightness on main sequence of Jl,0 = 6.20 (Dotter et al. 2008). The fiducial
lens distance is 5.3 kpc, hence, it should be behind almost all extinction observed in this field. The
expected brightness of the lens is hence Jl = 20.40 (see also Figure 4). The optical data and VVV
extinction predict the source brightness of Js = 14.66 ± 0.19. Hence, the lens is on the order of
6 mag fainter than the source and its contribution to the total light (0.004 mag) is much smaller
than the uncertainty in Js. The object brightness measured on the Keck image JKeck = 14.70± 0.03
Article number, page 12 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
is marginally brighter than the predicted source brightness Js = 14.66 ± 0.19, that is, no light from
the lens is detected. However, there is no statistical difference between the two measurements,
primarily due to the large uncertainty of J-band extinction.
5. Solving the mystery of OGLE-2002-BLG-045
Skowron et al. (2009) analyzed a number of microlensing events that were observed to repeat, that
is, show more than one brightening episode. The second episode can be caused by either a com-
panion to the lens or a companion to the source. The interesting finding by Skowron et al. (2009)
was that the event OGLE-2002-BLG-045 showed two consecutive OGLE datapoints that are well
separated from the main event and significantly brighter than the baseline. The two observations
happened only four days apart (HJD(cid:48) = 2455.7 and 2459.6) and were separated a few days from
the previous (HJD(cid:48) = 2448.7) and the following (HJD(cid:48) = 2463.6) observations, which both were
at the baseline. The short time between the two anomaly observations compared to tE = 26.4 d
suggests that the anomaly could have been caused by a planetary companion to the lens (q = 0.008
and s = 3.958). As was pointed out by Skowron et al. (2009), the only evidence for existence of
the planet were the two data points brighter than the baseline, and therefore, the planet detection of
the planetary companion was questionable. Because of this ambiguity, the putative planet OGLE-
2002-BLG-045Lb is not normally considered on the lists of the known microlensing planets (e.g.,
Zhu et al. 2014; Penny et al. 2016; Mróz et al. 2017).
In order to verify the planetary signal in OGLE-2002-BLG-045 we performed photometry of
the archival data acquired by the previous phase of the MOA survey (MOA-I; Yanagisawa et al.
2000; Bond et al. 2001). No signal of the planet was found. We also visually verified the OGLE
images that resulted in two anomalous points and found that they were taken in non-photometric
conditions. Skowron et al. (2009) inspected 4120 events, hence it is not surprising that in this
sample they found an event with two consecutive erroneous measurements separated by a few tE
from the event peak. We conclude that there is no convincing evidence that OGLE-2002-BLG-045L
has a wide separation planet and the two data points brighter than baseline are simply observational
artifacts.
6. Conclusions
We presented the discovery of MOA-2012-BLG-006Lb – an object a few times more massive than
Jupiter, which can be classified based on its mass either as a planet (most probable scenario) or a
brown dwarf (if its mass is at the high end of the derived distribution). We detected microlensing
signal not only due to this object but also due to its host star. The mass ratio is above 0.01, that
is, higher than the typical mass ratio of a protoplanetary disc to the parent star. Hence, the lower-
mass object could have formed independently and thus resembles brown dwarfs, even if its mass is
smaller than the commonly assumed boundary of 13 MJ.
Article number, page 13 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
The lack of parallax constraint precludes a direct measurement of the lens mass and projected
separation of its components, but the large ratio between the two subevents' time separation and
the event timescale suggests the companion is on a very wide orbit. Bayesian inference using a
Galactic model results in a 0.49± 0.25 M(cid:12) host orbited by a 8.4± 4.3 MJ companion at a projected
separation of 10.2 ± 2.2 a.u. The projected separation relative to Einstein ring radius of s = 4.4 is
the second largest among low-mass companions found by the microlensing technique after OGLE-
2008-BLG-092 (s = 5.3; Poleski et al. 2014) and preceding MOA-2007-BLG-400 (s = 2.9 if wider
of the two solutions is true; Dong et al. 2009).
The properties of MOA-2012-BLG-006Lb are similar to the small number of objects, either
high-mass planets or brown dwarfs, that have been discovered around M stars via direct imaging
and which typically have orbital separations of tens to hundreds of a.u. (see e.g., Table 1 of Lannier
et al. 2016). The masses of such planets tend to be at least the same order of magnitude as the total
amount of mass that comprised the protoplanetary disk within (and from) which we would expect
them to have formed, presenting a challenge to our current understanding of giant planet formation.
Nevertheless, the discovery of MOA-2012-BLG-006Lb suggests that whatever mechanisms are
responsible for the formation of such objects, they seem to operate similarly in the immediate
Solar neighborhood (where direct imaging finds them) and in other parts of the Galaxy, several kpc
away (where only microlensing is sensitive to their detection).
The AO image of the event was taken using the Keck NIRC-2 camera. The contribution of the
lens flux to the total observed flux could not be measured due to large uncertainty in extinction and
the fact that the lens is expected to be substantially fainter than the source.
We also showed that the anomaly observed in another event – OGLE-2002-BLG-045 – is of
instrumental origin. Hence, there is no planetary signature in that event.
Acknowledgements. The authors would like to thank Prof. A. Gould for consultation. The OGLE project has received fund-
ing from the National Science Centre, Poland, grant MAESTRO 2014/14/A/ST9/00121 to A.U. OGLE Team acknowledges
Profs. M. Kubiak and G. Pietrzy´nski, former members of the team, for their contribution to the collection of the OGLE
photometric data over the past years. J.P.B. and J.B.M. gratefully acknowledge support from ESO's DGDF 2014. C.C.
was supported by an appointment to the NASA Postdoctoral Program at NASA Ames Research Center, administered by
Universities Space Research Association under contract with NASA.
References
Alard, C. 2000, A&AS, 144, 363
Andrews, S. M., Rosenfeld, K. A., Kraus, A. L., & Wilner, D. J. 2013, ApJ, 771, 129
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2009, ApJ, 700, 1502
Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, A&A, 558, A33
Beaulieu, J.-P., Bennett, D. P., Batista, V., et al. 2016, ApJ, 824, 83
Beichman, C., Gelino, C. R., Kirkpatrick, J. D., et al. 2014, ApJ, 783, 68
Bennett, D. P., Bond, I. A., Udalski, A., et al. 2008, ApJ, 684, 663
Bensby, T., Adén, D., Meléndez, J., et al. 2011, A&A, 533, A134
Bertin, E. 2010, SWarp: Resampling and Co-adding FITS Images Together, Astrophysics Source Code Library
ascl.net/1010.068
Bertin, E. & Arnouts, S. 1996, A&AS, 117, 393
Bessell, M. S. & Brett, J. M. 1988, PASP, 100, 1134
Article number, page 14 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Bond, I. A., Abe, F., Dodd, R. J., et al. 2001, MNRAS, 327, 868
Boss, A. P., Basri, G., Kumar, S. S., et al. 2003, Brown Dwarfs – IAU Symposium, 211, 529
Cassan, A. 2008, A&A, 491, 587
Chabrier, G., Johansen, A., Janson, M., & Rafikov, R. 2014, Protostars and Planets VI, ArXiv e-prints 1401.7559, 619
Chauvin, G., Vigan, A., Bonnefoy, M., et al. 2015, A&A, 573, A127
Clanton, C. & Gaudi, B. S. 2014, ApJ, 791, 90
Claret, A. & Bloemen, S. 2011, A&A, 529, A75
Dodson-Robinson, S. E., Veras, D., Ford, E. B., & Beichman, C. A. 2009, ApJ, 707, 79
Dong, S., Bond, I. A., Gould, A., et al. 2009, ApJ, 698, 1826
Dong, S., Udalski, A., Gould, A., et al. 2007, ApJ, 664, 862
Dotter, A., Chaboyer, B., Jevremovi´c, D., et al. 2008, ApJS, 178, 89
Feroz, F. & Hobson, M. P. 2008, MNRAS, 384, 449
Feroz, F., Hobson, M. P., & Bridges, M. 2009, MNRAS, 398, 1601
Feroz, F., Hobson, M. P., Cameron, E., & Pettitt, A. N. 2013, ArXiv e-prints 1306.2144 [arXiv:1306.2144]
Foreman-Mackey, D., Morton, T. D., Hogg, D. W., Agol, E., & Schölkopf, B. 2016, AJ, 152, 206
Gaudi, B. S. 1998, ApJ, 506, 533
Gaudi, B. S. & Han, C. 2004, ApJ, 611, 528
Gonzalez, O. A., Rejkuba, M., Zoccali, M., et al. 2012, A&A, 543, A13
Gould, A. 2000, ApJ, 542, 785
Gould, A. 2008, ApJ, 681, 1593
Gould, A., Udalski, A., Shin, I.-G., et al. 2014, Science, 345, 46
Grether, D. & Lineweaver, C. H. 2006, ApJ, 640, 1051
Han, C. 2006, ApJ, 638, 1080
Jaroszynski, M. & Paczynski, B. 2002, Acta Astron., 52, 361
Kains, N., Browne, P., Horne, K., Hundertmark, M., & Cassan, A. 2012, MNRAS, 426, 2228
Kervella, P., Bersier, D., Mourard, D., et al. 2004, A&A, 428, 587
Lannier, J., Delorme, P., Lagrange, A. M., et al. 2016, A&A, 596, A83
Minniti, D., Lucas, P. W., Emerson, J. P., et al. 2010, New A, 15, 433
Mróz, P., Han, C., Udalski, A., et al. 2017, AJ, 153, 143
Nataf, D. M., Gould, A., Fouqué, P., et al. 2013, ApJ, 769, 88
Pejcha, O. & Heyrovský, D. 2009, ApJ, 690, 1772
Penny, M. T., Henderson, C. B., & Clanton, C. 2016, ApJ, 830, 150
Poleski, R., Skowron, J., Udalski, A., et al. 2014, ApJ, 795, 42
Ranc, C., Cassan, A., Albrow, M. D., et al. 2015, A&A, 580, A125
Sako, T., Sekiguchi, T., Sasaki, M., et al. 2008, Experimental Astronomy, 22, 51
Shvartzvald, Y., Maoz, D., Kaspi, S., et al. 2014, MNRAS, 439, 604
Shvartzvald, Y., Maoz, D., Udalski, A., et al. 2016, MNRAS, 457, 4089
Skowron, J. & Gould, A. 2012, ArXiv e-prints [arXiv:1203.1034]
Skowron, J., Udalski, A., Gould, A., et al. 2011, ApJ, 738, 87
Skowron, J., Udalski, A., Kozłowski, S., et al. 2016, Acta Astron., 66, 1
Skowron, J., Wyrzykowski, Ł., Mao, S., & Jaroszy´nski, M. 2009, MNRAS, 393, 999
Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163
Spiegel, D. S., Burrows, A., & Milsom, J. A. 2011, ApJ, 727, 57
Street, R. A., Choi, J.-Y., Tsapras, Y., et al. 2013, ApJ, 763, 67
Sumi, T., Bennett, D. P., Bond, I. A., et al. 2010, ApJ, 710, 1641
Sumi, T., Kamiya, K., Bennett, D. P., et al. 2011, Nature, 473, 349
Suzuki, D., Bennett, D. P., Sumi, T., et al. 2016, ApJ, 833, 145
Udalski, A. 2003, Acta Astron., 53, 291
Article number, page 15 of 16
Poleski et al.: Planet/brown dwarf on a wide orbit
Udalski, A., Szyma´nski, M. K., & Szyma´nski, G. 2015, Acta Astron., 65, 1
Wilson, P. A., Hébrard, G., Santos, N. C., et al. 2016, A&A, 588, A144
Wo´zniak, P. R. 2000, Acta Astron., 50, 421
Yanagisawa, T., Muraki, Y., Matsubara, Y., et al. 2000, Experimental Astronomy, 10, 519
Yoo, J., DePoy, D. L., Gal-Yam, A., et al. 2004, ApJ, 603, 139
Zhu, W., Penny, M., Mao, S., Gould, A., & Gendron, R. 2014, ApJ, 788, 73
1 Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA
e-mail: [email protected]
2 Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
3 Institute for Natural and Mathematical Sciences, Massey University, Private Bag 102904 North Shore Mail
Centre,
Auckland 0745, New Zealand
4 Sorbonne Universités, UPMC Univ Paris 6 et CNRS, UMR 7095, Institut d'Astrophysique de Paris, 98 bis
bd Arago, F-75014 Paris, France
5 LESIA Observatoire de Paris, Section de Meudon 5, place Jules Janssen F-92195 Meudon, France
6 School of Physical Sciences, University of Tasmania, Private Bag 37 Hobart, Tasmania 7001 Australia
7 NASA Ames Research Center, Space Science & Astrobiology Division, Moffett Field, CA 94035, USA
8 Department of Physics, University of Warwick, Coventry CV4 7AL, UK
9 Laboratory for Exoplanets and Stellar Astrophysics, NASA/Goddard Space Flight Center, Greenbelt, MD
20771, USA
10 Department of Earth and Space Science, Graduate School of Science, Osaka University, 1-1 Machikaneyama,
Toyonake, Osaka 560-0043, Japan
11 Department of Physics, University of Auckland, Private Bag 92019, Auckland, New Zealand
12 Institute of Space-Earth Environmental Research, Nagoya University, Furo-cho, Chikusa, Nagoya, Aichi
464-8601, Japan
13 Department of Physics, University of Notre Dame, Notre Dame, IN 46556, USA
14 Okayama Astrophysical National Astronomical Observatory, 3037-5 Honjo, Kamogata, Asakuchi, Okayama
719-0232, Japan
15 Nagano National College of Technology, Nagano 381-8550, Japan
16 Tokyo Metropolitan College of Industrial Technology, Tokyo 116-8523, Japan
17 School of Chemical and Physical Sciences, Victoria University, Wellington, New Zealand
18 University of Canterbury Mt John Observatory, PO Box 56, Lake Tekapo 7945, New Zealand
19 Department of Physics, Faculty of Science, Kyoto Sangyo University, 603-8555 Kyoto, Japan
Article number, page 16 of 16
|
1503.02581 | 1 | 1503 | 2015-03-09T17:48:33 | Chaos in navigation satellite orbits caused by the perturbed motion of the Moon | [
"astro-ph.EP"
] | Numerical simulations carried out over the past decade suggest that the orbits of the Global Navigation Satellite Systems are unstable, resulting in an apparent chaotic growth of the eccentricity. Here we show that the irregular and haphazard character of these orbits reflects a similar irregularity in the orbits of many celestial bodies in our Solar System. We find that secular resonances, involving linear combinations of the frequencies of nodal and apsidal precession and the rate of regression of lunar nodes, occur in profusion so that the phase space is threaded by a devious stochastic web. As in all cases in the Solar System, chaos ensues where resonances overlap. These results may be significant for the analysis of disposal strategies for the four constellations in this precarious region of space. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1–6 (2015)
Printed 10 March 2015
(MN LATEX style file v2.2)
Chaos in navigation satellite orbits caused by the perturbed motion
of the Moon
Aaron J. Rosengren,1⋆ Elisa Maria Alessi,1 Alessandro Rossi1
and Giovanni B. Valsecchi1,2
1IFAC-CNR, Via Madonna del Piano 10, 50019 Sesto Fiorentino (FI), Italy
2IAPS-INAF, Via Fosso del Cavaliere 100, 00133 Roma, Italy
ABSTRACT
Numerical simulations carried out over the past decade suggest that the orbits of the Global
Navigation Satellite Systems are unstable, resulting in an apparent chaotic growth of the ec-
centricity. Here we show that the irregular and haphazard character of these orbits reflects a
similar irregularity in the orbits of many celestial bodies in our Solar System. We find that
secular resonances, involving linear combinations of the frequencies of nodal and apsidal pre-
cession and the rate of regression of lunar nodes, occur in profusion so that the phase space is
threaded by a devious stochastic web. As in all cases in the Solar System, chaos ensues where
resonances overlap. These results may be significant for the analysis of disposal strategies for
the four constellations in this precarious region of space.
Key words: celestial mechanics – chaos – methods: analytical – methods: numerical – planets
and satellites: dynamical evolution and stability — planets and satellites: general.
1 INTRODUCTION
Space debris—remnants of past missions, satellite explosions, and
collisions—is a phenomenon that has existed since the beginning of
the space age; however, its significance for space activities, in par-
ticular the increasing impact risks posed to space systems, has been
realised only in the past few decades (Kessler & Cour-Palais 1978;
Rossi et al. 1999; Liou & Johnson 2006). The proliferation of space
debris has motivated deeper and more fundamental analysis of
the long-term evolution of orbits about Earth (Breiter 2001b,a;
Celletti & Gales¸ 2014). Orbital resonances are widespread within
this system as a whole (Hughes 1980), but particularly so amongst
the medium-Earth orbits (MEOs) of the navigation satellites in the
region of semimajor axes between 4 and 5 Earth radii, and a clear
picture of their nature is of great importance in assessing debris mit-
igation measures (Alessi et al. 2014). Indeed, the discovery that the
recommended graveyard orbits of these satellites, located several
hundred kilometres above the operational constellations, are poten-
tially unstable has led to a new paradigm in post-mission disposal—
one that seeks to cleverly exploit these dynamical instabilities
and the associated eccentricity growth for re-entry and destruction
within the Earth’s atmosphere (Jenkin & Gick 2002; Chao & Gick
2004; Rossi 2008; Deleflie et al. 2011). Previous studies have al-
ready noted the connection between the origin of the long-timescale
instabilities in the MEO region and a resonance phenomenon in-
volving Earth oblateness and lunisolar perturbations, yet very little
attention has been given to a true physical explanation of the erratic
⋆ E-mail: [email protected]
behaviour. To identify the source of orbital instability, we inves-
tigated the main resonant structures that organise and govern the
long-term orbital motion of navigation satellites. This paper clari-
fies the fundamental role played by secular resonances and reveals
the significance of the perturbed motion of the Moon.
2 ANALYSIS
for
the onset of dynamical
resonances are ubiquitous in the Solar System and
Orbital
instability and
are harbingers
chaos (Duncan et al. 1995; Lecar et al. 2001; Morbidelli 2002;
Colwell et al. 2009; Tsiganis 2010; Asphaug 2014; Lithwick & Wu
2014). Our knowledge of such phenomena in celestial mechanics
comes mainly from studies on asteroid dynamics—attempts to de-
scribe the stunningly complex dynamical structure of the asteroid
belt since the discovery of the Kirkwood gaps (Morbidelli 2002;
Tsiganis 2010). The orbital structure of the main belt is divided
by unstable regions, narrow gaps in the distribution of the aster-
oidal semimajor axes at 2.5 and 2.8 astronomical units (AU) from
the Sun, where the orbital periods (or mean motions) of the as-
teroids are commensurable to that of Jupiter. The outer boundary
is marked by the Jovian 2:1 mean-motion resonance at 3.3 AU.
Secular resonances, which involve commensurabilities amongst the
slow frequencies of orbital precession of an asteroid and the plan-
ets, define the inner edge of the main belt near 2.1 AU and demar-
cate the recognised subpopulations (asteroid families and groups).
Secular resonances also occur embedded within the libration zones
of the prominent Jovian mean-motion resonances, and it is pre-
2
A. J. Rosengren et al.
cisely the coupling of these two main types of resonance phenom-
ena that generates widespread chaos (Lecar et al. 2001; Morbidelli
2002; Tsiganis 2010). The creation of the Kirkwood gaps is thereby
associated with a chaotic growth of the asteroids’ eccentricities
to planet-crossing orbits, which leads to the removal from these
resonant regions by collisions or close encounters with the ter-
restrial planets. The same physical mechanism that sculpted the
asteroid belt—the overlapping of resonances—is responsible for
the emergence of chaos in the Solar System on all astronomical
scales, from the dynamics of the trans-Neptunian Kuiper belt and
scattered disk populations of small bodies (Duncan et al. 1995) to
the apparent perfectly regulated clockwork motion of the planets
(Lithwick & Wu 2014).
With the advent of artificial satellites and creation of space
debris, similar dynamical situations, every bit as varied and rich
as those in the asteroid belt, can occur in closer proximity to our
terrestrial abode (Upton et al. 1959; Hughes 1980; Breiter 2001a).
The dynamical environment occupied by these artificial celestial
bodies is subject to motions that are widely separated in frequency:
the earthly day, the lunar month, the solar year, and various preces-
sion frequencies ranging from a few years to nearly 26,000 years
for the equinoxes. A vast and hardly surveyable profusion of per-
turbations originates from the gravitational action of the Sun on the
Earth-Moon system. Primary among these irregularities is the re-
gression the Moon’s line of nodes with a period of about 18.61
years, and a progression of the lunar apsidal line with a period
of roughly 8.85 years. The provision of frequencies in the Earth-
Moon-Sun system gives rise to a diverse range of complex reso-
nant phenomena associated with orbital motions (Hughes 1980).
Indeed, past efforts to identify and classify the resonances that gen-
erate the orbital instability in the MEO region have been wholly ob-
scured by the abundance of frequencies in the orbital and rotational
motions of the Earth and of the Moon, and the great disparity of
timescales involved (Deleflie et al. 2011; Bordovitsyna et al. 2014;
Celletti & Gales¸ 2014).
While the orbits of most artificial satellites are too low to be
affected by mean-motion commensurabilities, even with the Earth’s
moon (Dichmann et al. 2014), there exists a possibility of a some-
what more exotic resonance involving a commensurability of a sec-
ular precession frequency with the mean motion of the Sun or the
Moon (Cook 1962; Hough 1981; Breiter 2001b). Among the more
compelling of these semi-secular commensurabilities is the evec-
tion resonance with the Sun, where the satellite’s apsidal preces-
sion rate produced by Earth’s oblateness equals the Sun’s apparent
mean motion, so that the line of apsides follows the Sun. This is
the strongest of the semi-secular apsidal resonances (Upton et al.
1959; Breiter 2001a), and its significance has even been stressed
in the dynamical history of the lunar orbit after its hypothesised
formation from an impact-generated disk (Asphaug 2014), as well
as in the dynamics of the Saturnian ring system with the larger,
more distant moons, Titan and Iapetus, playing the part of the
Sun (Colwell et al. 2009). It is therefore not surprising that some
authors (Deleflie et al. 2011; Bordovitsyna et al. 2014) have sug-
gested a possible link between these semi-secular resonances and
the irregular eccentricity growth in the orbits of the navigation
satellites. While it is true that such resonances exist for a wide
variety of eccentricities and inclinations, they are only of trivial
importance in these regions, as they mainly occur at much lower
semimajor axes (Breiter 2001a).
Satellite orbits resonant with respect to the perturbing influ-
ence of the Earth’s gravitational field are of particular importance
for the navigation satellite systems (Ely & Howell 1997; Ely 2002;
Rossi 2008; Celletti & Gales¸ 2014), and such resonance problems
have stood in the foremost rank of astrodynamical research work
(Breiter 2001a). Undoubtedly, the most celebrated resonance phe-
nomenon in artificial satellite theory is the critical inclination prob-
lem, which involves a commensurability between the two degrees
of freedom (apsidal and nodal motions) of the secular system
(Hough 1981; Delhaise & Morbidelli 1993; Tremaine & Yavetz
2014). The extreme proximity to these critical inclinations, espe-
cially in the case of the European Galileo and Russian GLONASS
constellations (Rossi 2008; Alessi et al. 2014), induced researchers,
very early, to associate the origin of the instabilities observed
in numerical surveys to these resonances (Jenkin & Gick 2002;
Chao & Gick 2004). However, these inclination-dependent-only
resonances are generally isolated (Cook 1962; Hughes 1980) and
thereby exhibit a well-behaved, pendulum-like motion, which,
when considered alone, cannot explain the noted chaotic be-
haviours. The second class of geopotential resonances arises from
the slight longitude-dependence of the geopotential, where the
satellite’s mean motion is commensurable with the rotation of the
Earth. Under such conditions, the longitudinal forces due to the
tesseral harmonics continually perturb the orbit in the same sense
and produce long-term changes in the semimajor axis and the mean
motion. For the inclined, slightly eccentric orbits of the naviga-
tion satellites, where these commensurabilities abound, the tesseral
harmonics exhibit a multiplet structure, akin to mean-motion reso-
nances, whereby the interaction of resonant harmonics can overlap
and produce chaotic motions of the semimajor axis (Ely & Howell
1997; Ely 2002). These effects, however, are localised to a nar-
row range of semimajor axis (tens of kilometres) and are of
much shorter period than secular precession (Ely & Howell 1997;
Celletti & Gales¸ 2014); consequently, tesseral resonances will not
significantly affect the long-term orbit evolution over timespans of
interest.
3 RESULTS AND DISCUSSION
These considerations have led us to investigate the role of lu-
nar secular resonances in producing chaos and instability among
the navigation satellites. We treat only the secular resonances of
the lunar origin, since, despite recent implications to the contrary
(Bordovitsyna et al. 2014), the solar counterparts will require much
longer timespans than what interests us here for their destabilising
effects to manifest themselves (Ely 2002)—a consequence of the
disparity of several orders of magnitude between the orbit preces-
sion periods of the Earth and the Moon. As a basis for our cal-
culations we have used a convenient trigonometric series develop-
ment by Hughes (1980), corresponding largely to a harmonic anal-
ysis of the perturbations. The lunar disturbing function is devel-
oped as a Fourier series of complicated structure, whose arguments
are combinations of the orbital phase and orientation angles of the
satellite and the Moon, and whose coefficients depend on the size
and shape (semimajor axes and eccentricities) of their orbits and
the inclinations. Two considerable simplifications are possible, re-
ducing this rather formidable expression in a marked degree. For
satellites whose semimajor axis does not exceed one tenth of the
Moon’s distance from the Earth, we can truncate the series to sec-
ond order in the ratio of semimajor axes, so that the lunar poten-
tial is approximated with sufficient accuracy by a quadrupole field
(Musen 1961; Delhaise & Morbidelli 1993; Rosengren & Scheeres
2013; Tremaine & Yavetz 2014). To study the secular interactions,
the short periodic terms of the disturbing function, depending on
the mean anomalies of both the satellite and the Moon, can be aver-
aged out (Morbidelli 2002). For lack of any mean-motion or semi-
secular resonances, such averaging effectively involves discarding
all terms which depend on these fast orbital phases. Accordingly,
the lunar disturbing function reduces to the form:
ω =
3
4
Ω = −
a (cid:17)2 5 cos2 i − 1
J2n(cid:16) R
(1 − e2)2 ,
a (cid:17)2
J2n(cid:16) R
(1 − e2)2 ,
cos i
3
2
(5)
Chaos in navigation satellite orbits
3
R =
µM a2
M (1 − e2
a3
M )3/2
2
2
X
m=0
X
s=0
(−1)mKm
(2 − s)!
(2 + m)!
2
F2,m,p(i)F2,s,1(iM )G2,p,2p−2(e)
(1)
p=0
X
(cid:2)(−1)2−sU m,−s
2
(ǫ) cos Φ+
p,m,s + U m,s
2
(ǫ) cos Φ−
p,m,s(cid:3) ,
with harmonic angles
Φ±
p,m,s = (2 − 2p)ω + mΩ ± sΩM ±
π
2
(s + m),
(2)
where m, s, and p are integers and the quantity Km is such that
K0 = 1 and Km = 2 for m > 0. The semimajor axis a, eccentric-
ity e, inclination i, longitude of ascending node Ω, and argument of
pericentre ω are the satellite’s orbital elements relative to the celes-
tial equator. The ecliptic orbital elements of the Moon are denoted
by subscript M, its gravitational mass by µM , and ǫ is the Earth’s
obliquity. The functions F and G are the modified Allan inclination
function and Hansen coefficient, respectively; and the function U
accounts for the fact that the Moon’s orbit is referred to the ecliptic
(see Hughes 1980, for more details):
U m,±s
2
=
(−1)m∓s
(2 ± s)!
(cos ǫ/2)m±s(sin ǫ/2)±s−m
d2±s
d Zn±s (cid:8)Z 2−m(Z − 1)n+m(cid:9) ,
(3)
where Z = cos2 ǫ/2. Such a choice of reference planes permits us
to consider iM as roughly constant and the variation in ΩM , largely
caused by the disturbing action of the Sun, as approximately linear
with period 18.61 years (Hughes 1980). The rather remarkable fact
that the Moon’s perigee does not explicitly appear in the quadrupo-
lar expansion of the secular problem has been known for some time
(Musen 1961), and belies the recent assertion (Bordovitsyna et al.
2014) that the position of the lunar perigee gives rise to resonance
effects on navigation satellite orbits.
A lunar secular resonance occurs when
ψ2−2p,m,±s = (2 − 2p) ω + m Ω ± s ΩM ≈ 0;
(4)
that is, when a specific linear combination of the secular precession
frequencies vanishes. For the Moon, the rate of change of ΩM is ap-
proximately −0.053 deg/d. The oblateness precession completely
overshadows the lunisolar effects and other perturbing influences
so that ω and Ω have an essentially linear time dependence,1 with
secular rates given by the classical expressions (Cook 1962):
where J2 is the second zonal harmonic coefficient of the geopo-
tential, R is the mean equatorial radius of the Earth, and n is the
satellite’s mean motion. Equations (4) and (5) define surfaces of
secular resonances in the space of the orbital elements a, e, and
i. As the semimajor axis is secularly invariable after averaging, the
commensurability condition defines curves of secular resonances in
eccentricity and inclination phase space. The interaction of various
lunar harmonics produces an exceedingly complicated network of
resonances—courtesy of the Earth’s oblateness—becoming partic-
ularly dense near the inclinations of the navigation satellite orbits
(Fig. 1). Each of the critical inclinations of the geopotential reso-
nances (corresponding to s = 02) split into a multiplet of resonant
curves, emanating from unity eccentricity.3 Each of the resonance
curves in Fig. 1 has its own pendulum-like structure, with a char-
acteristic domain of libration. When the separation between nearby
resonances becomes similar to their libration widths, the resonant
critical angles of the trajectories in these overlapping regions may
switch irregularly between libration and circulation; this alternation
is a hallmark of chaos (Morbidelli 2002; Lithwick & Wu 2014).
To gain an understanding of the physical significances of these
resonances, it it important to develop analytical models that ac-
curately reflect the true nature of the resonant interactions. The
simplest physical model for the study of the secular and resonant
motion in MEO consists of the quadrupole order of approximation
for the Earth, lunar, and solar gravitational potentials, but with the
Moon’s motion represented by a precessing and rotating elliptical
orbit. The singly-averaged model forms a nonautonomous Hamil-
tonian two degrees-of-freedom system, depending periodically on
time through the Moons perturbed motion (Rosengren & Scheeres
2013). Numerical simulations of this model show that the regions
where two or more neighbouring lunar secular resonances inter-
act exhibit chaos and large-scale excursions in eccentricity (Fig. 2),
in agreement with theoretical predictions based on the resonance
overlap criterion (Chirikov 1979). Chaotic diffusion is mediated by
the web-like structure of secular resonances, which permeate the
phase space and allows an initially circular orbit to become highly
eccentric, as revealed in a Poincar´e map of the nonautonomous,
periodically-perturbed two degrees-of-freedom system, obtained
by sampling the flow stroboscopically once per lunar nodal pe-
riod. Once a dynamical instability sets in, the subsequent evolu-
tion is highly chaotic and unpredictable in detail. The resonances
are the preferential routes for chaotic diffusion, as the trajectory
jumps from one resonance domain to the other. In any case, the
perigee will drive toward the ground, and the satellite is doomed
to destruction. The chaotic zones defined by the regions of overlap-
ping resonances do not preclude the existence of regular trajectories
embedded within it. Indeed, the character of the motion depends
sensitively upon the initial orientation angles of the satellite and
the initial lunar node (Fig. 2). The second orbit (bottom) suffers
1 This approximation is satisfactory when compared with the rigorous the-
ory. Neglect of the lunisolar perturbations on the the frequencies of nodal
and apsidal precession sets an upper limit to the radius of the orbit for which
the theory is valid; on the other hand, the period of precession must be ap-
preciably longer than a year for the double averaging procedure to be justi-
fiable, which sets a lower limit to the orbital radius. For these reasons, the
analysis is most useful in the region of semimajor axes between 4 and 5
Earth radii.
2 Hughes (1980) classifies these commensurabilities as inclination-
dependent-only lunisolar secular resonances, under the assumption that the
J2 harmonic in the geopotential produces the dominant change in the satel-
lite’s perigee and node (see also Cook 1962; Breiter 2001a).
3 To depict such curves up to e = 1 is only theoretical, as the satellites
will re-enter Earth’s atmosphere when e > 1 − R/a. Nevertheless, we are
mainly interested here in the early stages in the development of chaos.
4
A. J. Rosengren et al.
ψ2,1,±s
ψ2,0,±s
ψ 2,1,±s
ψ 2,2,±s
ψ0,m, s
0
50
100
150
200
250
300
350
400
450
500
y
t
i
c
i
r
t
n
e
c
c
E
1
0.8
0.6
0.4
0.2
0
0.8
0.6
0.4
0.2
0
30
40
50
60
70
80
Inclination (degrees)
Figure 2. Dynamical evolution of two Galileo-like orbits superimposed on
the background of complex resonant curves. The initial epoch of the sim-
ulations is 26 February 1998, which determines the initial dynamical con-
figuration of the Earth-Moon-Sun system, and the orbits were propagated
for 500 years (copper-tinge sequence of discrete points) using an accurate,
non-singular, singly-averaged model of oblateness and lunisolar perturba-
tions, develop previously by one of the authors (Rosengren & Scheeres
2013). Both representative Galileo orbits initially have a = 29, 600 km,
e = 0.001, i = 56◦, ω = 30◦, but their orbital planes are separated by
120◦ with Ω = 240◦ (top) and Ω = 120◦ (bottom), respectively. We con-
sider the discrete dynamics of the system by taking a subsequent snapshot
of the motion every lunar nodal period (a stroboscopic map), which gives
a faithful picture of the general character of the trajectories. The motions
with initial conditions in the overlapping regions can be both chaotic and
regular, depending on the initial orientation angles.
The structure of the whole phase space is far from being fully un-
derstood and clearly warrants greater theoretical development. Fu-
ture work will present more complicated maps, which allow for
the global visualisation of the geometrical organisation and coexis-
tence of chaotic and regular motion.
4 CONCLUSION
In the inner Solar System, overlapping secular resonances have
been identified as the origin of chaos in the orbits of the terres-
trial planets (Lecar et al. 2001; Morbidelli 2002; Lithwick & Wu
2014). We have demonstrated that the same underlying dynamical
mechanism responsible for the eventual destabilisation of Mercury,
and recently proposed to explain the orbital architecture of extra-
solar planetary systems (Lithwick & Wu 2014), is at the heart of the
orbital instabilities of seemingly more mundane celestial bodies—
the Earth’s navigation satellites. The occurrence and nature of the
secular resonances driving these dynamics depend chiefly on one
aspect of the Moon’s perturbed motion, namely, the regression of
the line of nodes. The decisive significance of this fact on space de-
bris mitigation will be emphasised in a later paper and its relation
to the design of disposal strategies in MEO (Alessi et al. 2014) will
be formulated there more precisely. The precarious state of these
a = 4.00R
ψ2,2,±s
1
0.8
0.6
0.4
0.2
0
0
10
20
30
40
50
60
70
80
90
y
t
i
c
i
r
t
n
e
c
c
E
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
a = 4.16R
0
10
20
30
40
50
60
70
80
90
a = 4.38R
0
10
20
30
40
50
60
70
80
90
a = 4.64R
0
10
20
40
60
30
Inclination (degrees)
50
70
80
90
Figure 1. The resonant structure of the medium-Earth orbits of the four
navigation constellations: GLONASS (a = 4.00R), GPS (a = 4.16R),
BeiDou (a = 4.38R), and Galileo (a = 4.64R). Each curve repre-
sents the location of the exact resonance of the form ψ2−2p,m,±s =
(2 − 2p) ω + m Ω ± s ΩM = 0, where the effects of perturbations on
ω and Ω other than the J2 harmonic have been neglected. For each group
of curves, the vertical line corresponds to the inclination-dependent-only
lunisolar secular resonances (s = 0), while the other curves are obtained
for s = −2, −1, 1, 2. Ely & Howell (1997) were the first to identify this
devious network of resonances for the GPS constellation; yet the exact na-
ture of its consequences has not been hitherto explored. These resonances
generally form the skeletal structure of the phase space, tracing the topolog-
ical organisation of the manifolds on which the chaotic motions take place
(Morbidelli 2002). Chaos is produced by the overlapping of the closely-
spaced resonant harmonics (Chirikov 1979).
only small-amplitude variations in eccentricity and inclination, and
is apparently regular, at least on the timescale of 500 years. Fig-
ure 3 shows that the absence of resonance overlapping generally
guarantees the local confinement of the motion. A large amount
of simulations, in which the initial perigee and nodes were varied,
were performed and further confirm this result: satellites with initial
conditions in the non-overlapping region exhibit regular, bounded
motions. For the full system one might expect that a coupling of
the tesseral and lunar resonance phenomena would produce a slow
diffusion in phase space, by which orbits can explore large regions
jumping from one resonance to another. A discussion of this aspect
of chaotic diffusion has been given by others (Ely & Howell 1997;
Ely 2002), and only occurs over millennia timescales.
It should be reiterated that the resonant curves presented in
Fig. 1 show the regions in the inclination-eccentricity phase space
for which chaotic orbits can be found and explain why chaotic or-
bits manifest in these regions only. It gives, however, no informa-
tion about which initial angles (ω, Ω, and ΩM ) will lead to chaos.
0
50
100
150
200
250
300
350
400
450
500
ACKNOWLEDGMENTS
Chaos in navigation satellite orbits
5
y
t
i
c
i
r
t
n
e
c
c
E
1
0.8
0.6
0.4
0.2
0
0.8
0.6
0.4
0.2
0
0
10
20
40
60
30
Inclination (degrees)
50
70
80
90
Figure 3. Dynamical evolution of five orbits with initial inclinations and
eccentricities in non-overlapping resonance regions superimposed on the
background of secular resonant curves. The initial epoch of the simu-
lations is 26 February 1998, and the orbits were propagated for 500
years (copper-tinge sequence of discrete points) using an accurate, non-
singular, singly-averaged model of oblateness and lunisolar perturbations
(Rosengren & Scheeres 2013). All orbits initially have a = 29, 600 km,
ω = 30◦, but their orbital planes are separated by 120◦ with Ω = 240◦
(top) and Ω = 120◦ (bottom), respectively. We consider the discrete dy-
namics of the system by taking a subsequent snapshot of the motion every
lunar nodal period (a stroboscopic map). The motions with initial conditions
in the non-overlapping regions are generally regular; chaotic motion can
only exist close to the separatrices of the isolated resonances (Morbidelli
2002).
constellations, perched on the threshold of instability, makes it un-
derstandable why all efforts to define stable graveyard orbits, espe-
cially in the case of Galileo, were bound to fail; the region is far too
complex to allow of an adoption of the simple geosynchronous dis-
posal strategy. A full understanding of the nature and consequences
of the chaos in these environments would have certainly helped in
the early design phases of the constellations.
this
represents,
Active debris removal, apart
in engineering that
from the daunting obsta-
cles
is currently seen
(Liou & Johnson 2006) as the only viable option to prevent the
self-generating Kessler syndrome phenomenon (collisional cascad-
ing; see Kessler & Cour-Palais 1978) from occurring in low-Earth
orbit, the most densely populated orbital environment (Rossi et al.
1999). But as we are still remarkably ignorant of the locations and
consequences of most resonances in near-Earth space, such drastic
measures may require a reassessment. This concerns particularly
the question as to whether strong instabilities exist, whose desta-
bilising effects occur on decadal timescales, that can be exploited
to effectively clear these regions of space from any future collision
hazard. Indeed, the process of dynamical clearing of resonant or-
bits is well illustrated by the paucity of asteroids observed in the
Kirkwood gaps (Tsiganis 2010).
The present form of the manuscript owes a great deal to the re-
viewer, Alessandra Celletti, of the University of Roma Tor Vergata;
by her various valuable suggestions, many concepts and results
have been clarified. We also owe special thanks to Florent Dele-
flie and J´erome Daquin, of the IMCCE/Observatoire de Paris, and
Kleomenis Tsiganis, of the Aristotle University of Thessaloniki,
for many insightful and motivating conversations. This work is
partially funded by the European Commissions Framework Pro-
gramme 7, through the Stardust Marie Curie Initial Training Net-
work, FP7-PEOPLE-2012-ITN, Grant Agreement 317185. Part of
this work was performed in the framework of the ESA Contract No.
4000107201/12/F/MOS “Disposal Strategies Analysis for MEO
Orbits”.
REFERENCES
Alessi E. M., Rossi A., Valsecchi G. B., Anselmo L., Pardini C.,
Colombo C., Lewis H. G., Daquin J., Deleflie F., Vasile M.,
Zuiani F., Merz K., 2014, Acta Astronaut., 99, 292
Asphaug E., 2014, Annu. Rev. Earth Planet. Sci., 42, 551
Bordovitsyna T. V., Tomilova I. V., Chuvashov I. N., 2014, Solar
Syst. Res., 48, 259
Breiter S., 2001a, Celest. Mech. Dyn. Astr., 81, 81
Breiter S., 2001b, Celest. Mech. Dyn. Astr., 80, 1
Celletti A., Gales¸ C., 2014, J. Nonlinear Sci., 24, 1231
Chao C. C., Gick R. A., 2004, Adv. Space Res., 34, 1221
Chirikov B. V., 1979, Phys. Rep., 52, 263
Colwell J. E., Nicholson P. D., Tiscareno M. S., Murray C. D.,
French R. G., Marouf E. A., 2009, in Dougherty M. K., Espos-
ito L. W., Krimigis S. M., eds, , Saturn from Cassini-Huygens.
Springer, Dordrecht, pp 413–458
Cook G. E., 1962, Geophys. J., 6, 271
Deleflie F., Rossi A., Portmann C., M´etris G., Barlier F., 2011,
Adv. Space Res., 47, 811
Delhaise F., Morbidelli A., 1993, Celest. Mech. Dyn. Astr., 57,
155
Dichmann D. J., Lebois R., Carrico Jr. J. P., 2014, J. Astronaut.
Sci., http://dx.doi.org/10.1007/s40295-014-0009-x
Duncan M. J., Levison H. F., Budd S. M., 1995, Astron. J., 110,
3073
Ely T. A., 2002, J. Guid. Cont. Dyn., 25, 352
Ely T. A., Howell K. C., 1997, Int. J. Dyn. Stab. Syst., 12, 243
Hough M. E., 1981, Celest. Mech., 25, 137
Hughes S., 1980, Proc. R. Soc. Lond. A, 372, 243
Jenkin A. B., Gick R. A., 2002, J. Spacecr. Rockets, 39, 532
Kessler D. J., Cour-Palais B. G., 1978, J. Geophys. Res., 83, 2637
Lecar M., Franklin F. A., Holman M. J., Murray N. W., 2001,
Annu. Rev. Astron. Astrophys., 39, 581
Liou J. C., Johnson N. L., 2006, Science, 311, 340
Lithwick Y., Wu Y., 2014, Proc. Natl. Acad. Sci. USA, 111, 12610
Morbidelli A., 2002, Modern Celestial Mechanics: Aspects of So-
lar System Dynamics. Taylor & Francis, London
Musen P., 1961, J. Geophys. Res., 66, 2797
Rosengren A. J., Scheeres D. J., 2013, Adv. Space Res., 52, 1545
Rossi A., 2008, Celest. Mech. Dyn. Astr., 100, 267
Rossi A., Valsecchi G. B., Farinella P., 1999, Nature, 399, 743
Tremaine S., Yavetz T. D., 2014, Am. J. Phys., 82, 769
Tsiganis K., 2010, Eur. Phys. J. Special Topics, 186, 67
Upton E., Bailie A., Musen P., 1959, Science, 130, 1710
6
A. J. Rosengren et al.
This paper has been typeset from a TEX/ LATEX file prepared by the
author.
|
1606.01785 | 1 | 1606 | 2016-06-06T15:28:54 | Predictions for Dusty Mass Loss from Asteroids during Close Encounters with Solar Probe Plus | [
"astro-ph.EP",
"astro-ph.SR"
] | The Solar Probe Plus (SPP) mission will explore the Sun's corona and innermost solar wind starting in 2018. The spacecraft will also come close to a number of Mercury-crossing asteroids with perihelia less than 0.3 AU. At small heliocentric distances, these objects may begin to lose mass, thus becoming "active asteroids" with comet-like comae or tails. This paper assembles a database of 97 known Mercury-crossing asteroids that may be encountered by SPP, and it presents estimates of their time-dependent visible-light fluxes and mass loss rates. Assuming a similar efficiency of sky background subtraction as was achieved by STEREO, we find that approximately 80% of these asteroids are bright enough to be observed by the Wide-field Imager for SPP (WISPR). A model of gas/dust mass loss from these asteroids is developed and calibrated against existing observations. This model is used to estimate the visible-light fluxes and spatial extents of spherical comae. Observable dust clouds occur only when the asteroids approach the Sun closer than 0.2 AU. The model predicts that during the primary SPP mission between 2018 and 2025, there should be 113 discrete events (for 24 unique asteroids) during which the modeled comae have angular sizes resolvable by WISPR. The largest of these correspond to asteroids 3200 Phaethon, 137924, 155140, and 289227, all with angular sizes of roughly 15 to 30 arcminutes. We note that the SPP trajectory may still change, but no matter the details there should still be multiple opportunities for fruitful asteroid observations. | astro-ph.EP | astro-ph | Earth, Moon, and Planets manuscript No.
(will be inserted by the editor)
Predictions for Dusty Mass Loss from Asteroids during Close
Encounters with Solar Probe Plus
Steven R. Cranmer
Received: 17 January 2016 / Accepted: 5 June 2016
Abstract The Solar Probe Plus (SPP) mission will explore the Sun's corona and innermost
solar wind starting in 2018. The spacecraft will also come close to a number of Mercury-
crossing asteroids with perihelia less than 0.3 AU. At small heliocentric distances, these
objects may begin to lose mass, thus becoming "active asteroids" with comet-like comae or
tails. This paper assembles a database of 97 known Mercury-crossing asteroids that may be
encountered by SPP, and it presents estimates of their time-dependent visible-light fluxes
and mass loss rates. Assuming a similar efficiency of sky background subtraction as was
achieved by STEREO, we find that approximately 80% of these asteroids are bright enough
to be observed by the Wide-field Imager for SPP (WISPR). A model of gas/dust mass loss
from these asteroids is developed and calibrated against existing observations. This model is
used to estimate the visible-light fluxes and spatial extents of spherical comae. Observable
dust clouds occur only when the asteroids approach the Sun closer than 0.2 AU. The model
predicts that during the primary SPP mission between 2018 and 2025, there should be 113
discrete events (for 24 unique asteroids) during which the modeled comae have angular sizes
resolvable by WISPR. The largest of these correspond to asteroids 3200 Phaethon, 137924,
155140, and 289227, all with angular sizes of roughly 15 to 30 arcminutes. We note that
the SPP trajectory may still change, but no matter the details there should still be multiple
opportunities for fruitful asteroid observations.
Keywords Asteroids · Comets · Inner Heliosphere
1 Introduction
Asteroids and comets are probes of the primordial solar system. Their weak gravitational
attraction enables the study of a range of physical processes that are not possible to de-
tect on larger moons and planets. In recent years, the traditional astronomical distinction
S. R. Cranmer
Department of Astrophysical and Planetary Sciences, Laboratory for Atmospheric and Space Physics, Uni-
versity of Colorado, Boulder, CO 80309, USA
Tel.: +1-303-735-1265
E-mail: [email protected]
6
1
0
2
n
u
J
6
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
8
7
1
0
.
6
0
6
1
:
v
i
X
r
a
2
S. R. Cranmer
between rocky asteroids and ice-rich comets has been replaced by the recognition of a con-
tinuous distribution in both composition and volatility (e.g., Weissman et al. 1989). Some
objects that were initially identified as asteroids have been seen to exhibit cometary out-
bursts (Hartmann et al. 1990; Mazzotta Epifani et al. 2011). On the other hand, some known
comets have become dormant as they apparently exhausted their volatile-rich outer layers
(Jenniskens 2008; Ye et al. 2016). Recent work on active asteroids (Jewitt 2012; Jewitt et al.
2013, 2015; Agarwal et al. 2016) has shown that dusty mass loss may occur even when vir-
tually no icy material is left.
When small solid bodies approach the Sun, there are strong radiative and thermal effects
that release gas molecules, dust particles, and larger pieces of regolith (e.g., Delbo et al.
2014). At small heliocentric distances-e.g., for sungrazing comets-the dust dissociates
rapidly and the gas becomes ionized (Povich et al. 2003; Bryans and Pesnell 2012) and of-
ten the entire object is destroyed (Biesecker et al. 2002). Mass loss from comets in the in-
ner heliosphere remains a useful, albeit indirect, probe of the solar wind (Brandt and Snow
2000; Huebner et al. 2007) and the hot solar corona (Raymond et al. 2014). Many similar
ablative processes may also be occurring in the environments of extrasolar planets that orbit
close to their host stars (Mura et al. 2011; Matsakos et al. 2015).
Active asteroids in the innermost heliosphere have not yet been explored by plane-
tary spacecraft. However, Solar Probe Plus (SPP) will spend several years inside the or-
bit of Venus (McComas et al. 2007; Fox et al. 2015) with a minimum perihelion distance
of 0.0459 AU (i.e., 9.86 solar radii). In addition to a suite of in situ plasma and field in-
struments, the the Wide-field Imager for SPP (WISPR, Vourlidas et al. 2015) will observe
visible-light photons over large fields of view. The primary goal of WISPR is to observe K-
corona emission from Thomson-scattered electrons and F-corona emission from dust, but it
will also search for sungrazing comets and putative Vulcanoids (see, e.g., Steffl et al. 2013).
This paper explores the ability of instruments such as WISPR to observe extended emis-
sion from mass-losing active asteroids in the inner heliosphere. Sect. 2 of this paper surveys
the orbital properties of 97 Mercury-crossing asteroids, in both the Apollo (a > 1 AU) and
Aten (a < 1 AU) groups, that could be encountered by SPP. The sizes and visible-light fluxes
of these asteroids are estimated in Sect. 3 and compared with expected background levels of
zodiacal light. Sect. 4 presents a model for the mass loss rate of high latent-heat silicate ma-
terial from the selected asteroids, and Sect. 5 estimates the observable spatial extent of dusty
coma/tail regions that WISPR can resolve. Lastly, Sect. 6 discusses some of the the broader
implications of this work and gives suggestions for future improvements in the modeling.
2 Orbital Analysis
Figure 1 illustrates the latest version of the proposed SPP baseline mission trajectory. This
information was extracted from a SPICE kernel file distributed to the SPP team in Septem-
ber 2014. It assumes a launch date of July 31, 2018, and it extends to September 1, 2025.
The kernel data were processed with version N65 of the SPICE toolkit for IDL (Acton
1996), which was created by the NASA/JPL Navigation and Ancillary Information Facility
(NAIF).1 The Cartesian (x, y, z) positions of SPP, in solar-system barycenter coordinates for
epoch J2000.0, were computed and saved at 0.1 day intervals. The mission time t shown in
Figure 1(b) is specified in days measured from 0:00 UT on July 31, 2018.
1 http://naif.jpl.nasa.gov/naif/
Predictions for Dusty Mass Loss from Asteroids
3
Fig. 1 (a) Planned SPP trajectory shown projected in the ecliptic plane, compared with orbits of Venus and
the Earth (gray tracks). (b) Heliocentric radial distance of SPP shown versus mission time in days. In both
panels, the position of the Earth on July 31, 2018 is indicated with a black circle, and mission time is mapped
to curve color.
The dynamical properties of 97 Mercury-crossing asteroids were obtained from the
NASA/JPL Horizons On-Line Ephemeris System (Giorgini et al. 1996; Giorgini 2011)2 in
July 2015. The Horizons system was used to convert the tabulated orbital elements into
Cartesian (x, y, z) positions, which we saved at 12 hour intervals during the calendar years
2018 to 2026. In order to compare directly with the SPP trajectory data, each asteroid's co-
ordinates were interpolated to the denser (0.1 day) time grid obtained from the spacecraft
SPICE kernel.
The specific asteroids appropriate for this study were selected with the following criteria.
First, we included all 63 known asteroids with perihelia less than 0.2 AU. The asteroid that
approaches closest to the Sun is 2005 HC4, with perihelion q = 0.07066 AU (i.e., 15.2 R⊙).
2 http://ssd.jpl.nasa.gov/horizons.cgi
4
S. R. Cranmer
There are six others with perihelia less than 0.1 AU. Second, the database was extended to
asteroids with perihelia between 0.2 and 0.3 AU. However, out of the 159 known asteroids
in this region, many of them are extremely small and dim. The results of Sections 4–5 below
show that only the largest asteroids at q > 0.2 AU are expected to display substantial mass
loss. Thus, we selected the brightest 34 of that group, with the criterion that their V -band
absolute magnitudes H must be ≤ 18.0. No asteroids were chosen with q > 0.3 AU because
of both their negligible expected mass loss rates and the infrequency of close encounters
with SPP once it enters the inner heliosphere. Table 1 gives the final list of 97 asteroids in
order of increasing perihelion q, and it also lists their names/numbers, eccentricities e, and
V -band magnitudes H.
A set of 97 timelines containing the relative positions of SPP, the asteroid, and the Sun
were produced from the ephemeris data over the 2018–2025 mission period. The three mu-
tual distances d (between SPP and asteroid), ra (between Sun and asteroid), and rp (between
Sun and SPP) are used for various purposes in the models described below. Other useful
quantities include the solar elongation angle e (i.e., the angle centered on SPP between vec-
tors pointing to the Sun and to the asteroid) and the scattering phase angle a
(i.e., the angle
centered on the asteroid between vectors pointing to the Sun and to the observer on SPP).
Both angles can be computed from the three distances,
cose =
r2
p + d2 − r2
a
2rpd
cosa =
,
r2
a + d2 − r2
p
2rad
.
(1)
Table 1 lists dmin, the distance of closest approach for each asteroid to SPP, and the value of
ra at that time. The asteroid with the smallest value of dmin is 2007 EB26, with a minimum
separation of only 0.0137 AU (i.e., 2.95 R⊙), which should occur at t = 779.5 days. There
are 24 asteroids in the list that come closer than 0.1 AU to SPP.
Figure 2 shows the spread of minimum distance dmin versus each asteroid's perihelion
distance q. There is no overall correlation between these quantities, but there does appear to
be a slight preponderance (see dashed line) for asteroids with the largest q values to avoid
close approaches with SPP. This is likely to be a statistical trend associated with the fact
that asteroids with wider orbits naturally spend less time in SPP's neighborhood close to the
Sun.
It should be emphasized that the accurate prediction of a close approach between SPP
and any specific asteroid depends on the validity of the planned trajectory and launch date
of July 31, 2018. Spacecraft launches are frequently delayed, but SPP does have firm re-
quirements to "meet" the desired gravitational assists with Venus. Thus, it is possible that
even a delayed launch could result in an eventual synchronization with the trajectory as-
sumed here. In any case, some results of this paper may be better interpreted as one possible
sample from a quasi-random distribution of possible SPP orbits. Additional statistical con-
clusions are discussed in Section 6.
As an example of an interesting encounter, Figure 3(a) shows the mutual trajectories for
SPP and asteroid 137924 (2000 BD19). This example is only the 19th closest approach out
of the full list of 97 asteroids, but it is notable for occurring very near the asteroid's own
perihelion. That fact leads to it having a bright V magnitude and unusually high predicted
values for its mass loss rate and coma size (see below). Figure 3(b) also shows that, for the
case of asteroid 137924, multiple close encounters with SPP tend to occur near successive
perihelion passes.
Predictions for Dusty Mass Loss from Asteroids
5
Table 1 Inner heliospheric asteroids tracked in this study, sorted by perihelion.
Number
Name
- 2005 HC4
- 2008 FF5
- 2015 EV
394130
137924
374158
394392
2006 HY51
2000 BD19
2004 UL
2007 EP88
- 2011 KE
- 2008 HW1
- 2015 HG
- 2012 US68
- 2011 XA3
2002 PD43
2008 XM
2008 HE
399457
386454
431760
- 2007 EB26
2002 AJ129
276033
425755
289227
- 2000 LK
2011 CP4
- 1995 CR
- 2007 GT3
- 2004 QX2
- 2011 BT59
2004 XY60
- 2015 KO120
- 2007 PR10
- 2006 TC
- 2013 JA36
- 2008 MG1
- 2013 HK11
3200
Phaethon
- 2013 YC
- 2010 JG87
- 2015 KP157
- 2015 DU180
- 2012 UA34
- 2005 EL70
155140
364136
105140
302169
141851
267223
259221
2005 UD
2006 CJ
2000 NL10
- 2011 WN15
- 2013 WM
2001 TD45
- 2005 RV24
- 2008 EY68
2002 PM6
2001 DQ8
- 2013 AJ91
2003 BA21
- 2011 YX62
1566
89958
Icarus
2002 LY45
- 2009 HU58
5786
Talos
- 2003 UW29
q [AU]
0.07066
0.07914
0.08003
0.08100
0.09199
0.09283
0.09558
0.10013
0.10171
0.10472
0.10566
0.10859
0.11031
0.11107
0.11337
0.11573
0.11671
0.11788
0.11813
0.11931
0.12088
0.12498
0.12859
0.13017
0.13120
0.13241
0.13561
0.13750
0.13886
0.13901
0.14004
0.14104
0.14432
0.14820
0.15228
0.15597
0.15893
0.16287
0.16580
0.16727
0.17285
0.17466
0.17733
0.17805
0.17888
0.17955
0.18138
0.18187
0.18350
0.18409
0.18652
0.18675
0.18686
0.18727
0.18899
e
H [mag]
dmin [AU]
ra(dmin) [AU]
0.96119
0.96539
0.96100
0.96884
0.89505
0.92670
0.88584
0.95502
0.96061
0.95025
0.95776
0.92597
0.95603
0.90913
0.94993
0.78867
0.91488
0.94590
0.87039
0.86846
0.93938
0.90291
0.94848
0.79669
0.92577
0.89262
0.91184
0.94854
0.82271
0.93678
0.88984
0.94347
0.94773
0.91027
0.92097
0.80155
0.94022
0.87224
0.75492
0.81704
0.85793
0.91598
0.77742
0.88177
0.75994
0.85012
0.90150
0.92818
0.83321
0.92823
0.82696
0.88625
0.90955
0.82684
0.83840
20.7
23.1
22.5
17.2
17.2
18.8
18.5
19.8
17.4
21.0
18.3
20.5
19.1
20.0
18.1
19.6
18.7
18.4
21.2
21.7
19.7
21.7
21.0
18.9
22.0
20.7
18.8
21.0
19.9
20.7
14.6
21.3
19.1
19.2
20.8
19.5
24.0
17.3
20.2
15.8
19.6
23.8
19.9
20.6
22.0
17.7
18.0
19.3
19.1
23.0
16.9
17.0
19.1
17.1
20.7
0.10941
0.28942
0.64377
0.32687
0.07756
0.16533
0.12213
0.15917
0.55013
0.35597
0.15636
0.26899
0.41708
0.13714
0.21685
0.01372
0.36247
0.54865
0.12991
0.20231
0.55496
0.07061
0.25844
0.07934
0.05460
0.21100
0.22618
0.43394
0.03179
0.04258
0.24292
0.63135
0.48764
0.17604
0.24979
0.05960
0.10759
0.06729
0.02326
0.23609
0.17488
0.53346
0.12274
0.20295
0.12325
0.08396
0.06021
0.18173
0.06899
0.26575
0.21870
0.14820
0.11882
0.32613
0.06178
0.11916
0.46209
0.07659
0.41537
0.10245
0.22027
0.21280
0.39329
0.25181
0.13877
0.46061
0.39000
0.48005
0.29387
0.12394
0.20750
0.15895
0.26888
0.24157
0.23316
0.12185
0.69866
0.14597
0.13191
0.67427
0.75243
0.23233
0.58412
0.90056
0.15407
0.26831
0.15442
0.40593
0.57849
0.18051
0.35570
0.36308
0.17065
0.21205
0.44102
0.17701
0.25236
0.59688
0.75445
0.17417
0.58174
0.18495
0.32416
0.23841
0.46486
0.29748
0.22125
0.34948
0.78989
0.35127
6
S. R. Cranmer
Table 1 (continued)
Inner heliospheric asteroids tracked in this study, sorted by perihelion.
Number
Name
387505
1998 KN3
- 2007 MK6
- 2015 KJ122
- 2015 DZ53
- 2010 VA12
- 1996 BT
153201
139289
66391
141079
143637
329915
438116
369296
184990
2000 WO107
2001 KR1
1999 KW4
2001 XS30
2003 LP6
2005 MB
2005 NX44
2009 SU19
2006 KE89
- 2004 LG
- 2005 GL9
137052
225416
242643
136874
40267
Tjelvar
1999 YC
- 2000 SG8
2005 NZ6
1998 FH74
1999 GJ4
- 2011 WS2
- 2007 VL243
1984 QY1
- 2006 OS9
2010 LG14
2004 VA64
2005 EY
2004 EC
2002 WZ2
2006 KZ112
2002 UR3
2008 EM9
- 2014 MR26
2001 FO32
1989 VA
2003 WM7
331471
369452
190119
351370
164201
385402
397237
253106
364877
231937
99907
170502
- 2010 KY127
162269
141525
1999 VO6
2002 FV5
q [AU]
0.19537
0.19586
0.19613
0.19620
0.19875
0.19978
0.19985
0.19996
0.20010
0.20015
0.20341
0.20411
0.20495
0.20935
0.21144
0.21250
0.22226
0.23768
0.24099
0.24508
0.24872
0.25390
0.25669
0.25890
0.26200
0.26348
0.26379
0.26898
0.27010
0.27510
0.28058
0.28476
0.28545
0.28549
0.29101
0.29387
0.29523
0.29525
0.29648
0.29686
0.29734
0.29916
e
H [mag]
dmin [AU]
ra(dmin) [AU]
0.87328
0.81879
0.75026
0.87013
0.84334
0.83500
0.78072
0.84123
0.68846
0.82815
0.88352
0.79284
0.90745
0.89942
0.79925
0.89714
0.89620
0.80955
0.83050
0.90066
0.86443
0.88462
0.80825
0.74356
0.72856
0.89453
0.90377
0.74267
0.89042
0.89066
0.85954
0.88432
0.88694
0.79295
0.85153
0.76593
0.82644
0.59468
0.88027
0.88116
0.73809
0.72475
18.4
19.9
22.0
20.8
19.5
23.0
19.3
17.6
16.5
17.7
16.3
17.1
17.3
17.9
16.4
18.0
17.1
16.9
17.2
17.5
17.4
15.7
15.4
17.2
17.8
15.4
17.8
17.9
17.1
17.2
15.7
17.0
16.7
16.4
17.3
17.8
17.7
17.9
17.2
17.0
17.0
17.9
0.18166
0.29098
0.06455
0.40446
0.07376
0.10126
0.19124
0.02376
0.12587
0.03979
0.31278
0.33105
0.35076
0.06908
0.14092
0.39704
0.03307
0.08579
0.29616
0.59763
0.08031
0.29403
0.31560
0.16082
0.04606
0.28198
0.21360
0.32184
0.41024
0.12769
0.12414
0.47899
0.75831
0.52059
0.08666
0.12435
0.20666
0.24233
0.20954
0.51914
0.26966
0.21795
0.20196
0.37447
0.50995
0.20443
0.22421
0.37595
0.35397
0.47075
0.20703
0.20640
0.51759
0.54317
0.81328
0.26797
0.52566
0.38711
0.22339
0.24021
0.31271
0.75514
0.81175
0.29322
0.25078
0.32279
0.46376
0.34163
0.35270
0.36409
0.37468
0.27657
0.28329
0.34397
0.66348
0.66113
0.34927
0.66948
0.32020
0.31057
0.77615
0.36652
0.34300
0.33586
3 Asteroid Physical Properties
In the analysis performed below, each asteroid is assumed to be spherical in shape, with a
diameter D given by a standard conversion from its visible-light absolute magnitude H,
D = (cid:16)1348 p−1/2 10−H/5(cid:17) km ,
(2)
(e.g., Bowell et al. 1989), where we take p = 0.1 as a representative value of the geo-
metric albedo for small asteroids in the inner heliosphere (see also Muinonen et al. 1995;
Morbidelli et al. 2002). For the 97 asteroids listed in Table 1, the median derived value of
Predictions for Dusty Mass Loss from Asteroids
7
Fig. 2 Minimum distances dmin between SPP and each asteroid, found by comparing ephemerides between
July 31, 2018 and September 1, 2025. The symbol colors map to asteroid absolute magnitude H.
D is 0.890 km, and the minimum and maximum values are 0.0675 km (2005 EL70) and
5.12 km (3200 Phaethon), respectively. Their angular sizes, as seen by SPP, are typically in
the range between 0.001′′ and 0.03′′, with the largest value of 0.075′′ found for the closest
approach of asteroid 2001 KR1. These values are much smaller than the spatial resolution
of the WISPR telescopes. However, we anticipate that many asteroids will emit bright dust
clouds that extend to distances several orders of magnitude larger than their respective di-
ameters (see Section 5).
The apparent magnitude of an asteroid in the Johnson–Cousins V band (dominated by
wavelengths between 500 and 600 nm) is given by
F (a )
mV = H + 5log10(rad) − 2.5log10
(3)
where ra and d are given in units of AU. F (a ) is the phase function for the scattering
of sunlight, which is largest at opposition (a = 0) and decreases monotonically for larger
scattering angles. The standard phase function defined by Bowell et al. (1989) was used
with a slope parameter G = 0.15 appropriate for inner heliospheric asteroids (see also
Lagerkvist and Magnusson 1990). Apparent magnitudes were converted into V -band energy
fluxes at l ≈ 500 nm, with
F = 3.67 × 10−23 (10−0.4mV ) W m−2 Hz−1
(4)
using the normalization factor specified by Wamsteker (1981). Figure 4(a) shows the dis-
tribution of mV versus minimum distance dmin. The brightest one is asteroid 141079, with
mV = 8.95, but there is nothing unusual about its intrinsic properties. Finding a small value
of mV depends on the chance of finding small values for d, ra, and a
, all occurring at roughly
the same time.
The observability of a given asteroid depends on both its intrinsic brightness and its rel-
ative contrast with the sky background. Jewitt et al. (2013) observed 3200 Phaethon and its
8
S. R. Cranmer
Fig. 3 (a) Trajectories for the closest encounter between SPP (black curve) and asteroid 137924 (red curve).
Large symbols show positions at minimum distance, at mission time 1493.68 days. Smaller symbols show
the positions along a sequence of 7.2 hr intervals before and after the time of minimum distance. (b) Relative
distance between SPP and asteroid 137924 versus mission time in days. Red arrows show the times of the
asteroid's perihelia.
tail with the Sun Earth Connection Coronal and Heliospheric Investigation (SECCHI) pack-
age on STEREO (Howard et al. 2008; Eyles et al. 2009), and they noted how measurements
were limited by the presence of extended sky emission. The WISPR instrument on SPP will
be closer to its asteroid targets than was SECCHI, and it will have comparable sensitivity
despite its smaller size. Thus, in this paper we apply several of the lessons learned from
SECCHI to future measurements with WISPR.
Visible-light sky emission in the inner heliosphere is dominated by a combination of
zodiacal light (the dust-scattered F corona) and Thomson-scattered electron emission (the K
corona). A relatively simple analytic expression for the total specific intensity BZ(e , rp) of
the two components was found to reproduce a range of observations and model predictions.
Predictions for Dusty Mass Loss from Asteroids
For a standard observer at rp = 1 AU, the angular dependence is given by
BZ(e , 1 AU)
B⊙
=
4 × 10−25
e 8
+
6.3 × 10−14
[sin(0.63e )]2.22
9
(5)
where the elongation angle e
is specified in radians and the mean solar-disk brightness is
B⊙ = 2.96 × 10−8 W m−2 sr−1 Hz−1 in the V band (Allen 1973). Equation (5) matches col-
lected observations in the ecliptic plane (Leinert 1975; Munro & Jackson 1977; Kwon et al.
2004; Mann et al. 2004) to within about a factor of two. Observers closer to the Sun are
expected to see higher intensities (e.g., van Dijk et al. 1988), and we use
BZ(e , rp) = BZ(e , 1 AU)(cid:18) 1 AU
rp (cid:19)2.5
(6)
where the exponent 2.5 was estimated from the modeled inner heliospheric intensities in
Figure 7 of Vourlidas et al. (2015).
The goal is to compare the above sky brightness with the flux from an asteroid, but the
latter is essentially a tiny point-source with an angular size much smaller than a WISPR
detector pixel. The asteroid's flux F must be compared with a corresponding sky flux FZ
that fills the pixel. The zodiacal light intensity BZ is thus converted into flux by multiplying
it by the solid angle subtended by a pixel, with
FZ = W BZ(e , rp) .
(7)
Vourlidas et al. (2015) specified pixel sizes of 1.2′ and 1.7′ for the inner and outer fields of
view, and we used the average of the two as a representative value. Thus, a square patch
of the sky subtending 1.45′ = 4.22 × 10−4 radians on a side occupies a solid angle W =
1.78 × 10−7 sr.
Figure 4(b) shows the ratio F/FZ as a function of dmin. Differences with panel (a) are
mainly due to the fact that different asteroids are viewed with different elongation angles,
so the sky brightness varies. The highest-contrast asteroid is 2001 KR1, with F/FZ = 0.72.
Ideally, one would consider a "good" observation to be one with F ≫ FZ, but experience
with SECCHI has shown that much weaker signals can be extracted from strong back-
grounds. DeForest et al. (2011) used sophisticated processing techniques to resolve features
with fluxes as low as (F/FZ ) ≈ 10−4. That level is indicated with a horizontal line in Fig-
ure 4(b), and we note that 76 out of 97 asteroids fall above that level. WISPR will clearly
have multiple opportunities to observe asteroids in the inner heliosphere.
It is noteworthy that the list of five brightest (i.e., lowest mV ) asteroids and the list of
five highest flux contrast (largest F/FZ) asteroids share three members: 141079, 2005 GL9,
and 2007 EB26. Note that asteroid 137924, whose orbit is shown in Figure 3, is the second
brightest (mV = 8.98) at closest approach, but due to a low elongation angle it is only the
14th highest flux contrast (F/FZ = 0.038). The largest active asteroid, 3200 Phaethon, has
its lowest value of mV = 10.53 at t = 1384.6, about six days prior to the time it reaches its
minimum distance dmin to SPP. However, because of varying elongation angles, Phaethon's
time of maximum flux contrast (at F/FZ = 0.13) occurs at its previous perihelion passage
17 months earlier (t = 880.3). This tells us that the "snapshots" of the 97 closest-approach
events, illustrated by filled circles in Figures 2 and 4, may not always be the most reliable
guide to identifying when interesting things are occurring.
10
S. R. Cranmer
Fig. 4 (a) Apparent magnitudes, at times of closest approach with SPP, plotted versus minimum distance
between each asteroid and SPP. (b) V -band asteroid flux F divided by the estimated zodiacal light flux FZ
measured in a single WISPR pixel. The horizontal line shows an approximate observable threshold of 10−4
times the background. In both panels, symbol colors correspond to H with the same scaling as in Figure 2.
4 Asteroid Mass Loss due to Erosion
In a similar manner to comets, it is believed that active asteroids lose mass when they orbit
sufficiently close to the Sun. The ejected gas and dust expands to fill coma-like or tail-like
atmospheres that may be observed at large distances from the nucleus. One can use the lan-
guage of sublimation to discuss the erosion of solid material from the asteroid surface. Some
active asteroids (e.g., 133P/Elst-Pizzaro in the main asteroid belt; see Hsieh et al. 2004) are
probably similar to comets in that they emit gaseous mass mostly in the form of volatile
ices (H2O, CO, CO2). However, the near-Sun active asteroids (e.g., Phaethon) have likely
already lost most of these easily sublimated compounds. The heavier silicate and hydrocar-
bon molecules that presumably dominate the outer regolith layers of these asteroids have a
Predictions for Dusty Mass Loss from Asteroids
11
substantially higher latent heat than volatile comet ices. This section applies the long history
of sublimation energy-balance models for comets (e.g, Weigert 1959; Delsemme and Miller
1971; Whipple and Huebner 1976; Weissman 1980; Prialnik et al. 2004) to the case of near-
Sun active asteroids.
Before proceeding with such modeling, it is important to note that the loss of both
silicate-rich gas and dust from rocky objects (heated to T ≈ 1000 K) has not been stud-
ied as extensively as the standard cometary scenario of sublimating ice molecules that drag
along the larger dust grains. However, the idea of a simultaneous ejection of multiple phases
of the same type of material has been studied for comets. Water ice and other volatile com-
pounds appear to be ejected in both the gas phase and in the form of "snow" particles with
sizes ranging from microns to meters (e.g., Delsemme and Miller 1971; Kelley et al. 2013,
2015; Protopapa et al. 2014).
In addition to the analogy with comet ice loss, there are three other comparable situations
that appear to have some resemblance and relevance to the case of active asteroids:
1. Meteors entering a planetary atmosphere undergo rapid deceleration and thermal abla-
tion (e.g., Baldwin and Sheaffer 1974). Although the source of heating for meteors is
different (atmospheric drag versus solar radiation), the mass loss is often treated by re-
placing the latent heat of sublimation with a comparable heat of ablation (Chyba et al.
1993). The size distribution of resulting fragments appears to be quite broad, from
nanometer-scale "smoke" to micron-scale dust to macroscopic meteoroids (Borovicka and Charv´at
2009; Malhotra and Mathews 2011).
2. Sungrazing comets have perihelia within a few solar radii of the Sun's surface (Marsden
2005), and observations of their evolution upon close approach provide valuable infor-
mation about the composition and thermal properties of primordial bodies in the so-
lar system. Sekanina (2003) was able to model the light curves of a number of sun-
grazing comets by assuming a continuous distribution of silicate erosion products (i.e.,
from large fragments to individual sublimated molecules). These comets have also been
observed to emit metallic-i.e., alkali, sulfide-rich, and iron-group-material as well
(Preston 1967; Slaughter 1969; Zolensky et al. 2006; Ciaravella et al. 2010),
3. We now know that a large number of extrasolar planets orbit within just a few stel-
lar radii of their host stars (Winn and Fabrycky 2015). Some observations suggest that
small, rocky planets of this kind are slowly disintegrating via the ejection of large
amounts of dust (e.g., Mura et al. 2011; Rappaport et al. 2012; van Lieshout et al. 2014;
Sanchis-Ojeda et al. 2015). Possible formation channels for the dust include condensa-
tion from sublimated gas, direct ejection via volcanism, or comet-like entrainment of
grains along with escaping gas molecules.
Thus, there appear to be multiple ways that silicate-rich gas can be removed from the sur-
faces of solid objects near the Sun, together with larger dust grains, when they experience
strong solar irradiation. It should also be mentioned that the minerals left on the surfaces of
these bodies may be irradiated sufficiently to induce various types of chemical and tensile
metamorphosis (Scott et al. 1989; Scheeres 2005; Kasuga et al. 2006; Gundlach and Blum
2016).
The thermal energy balance at the surface of a solid body can be solved to compute the
mass loss rate M and equilibrium temperature T of molecules leaving the surface. Ignoring
heat conduction, the short-wave radiative energy gained must be balanced by losses due to
long-wave radiation and sublimation, with
(1 − A)S⊙
r2
a
cosq = hs
BT 4 + muLZ
(8)
12
S. R. Cranmer
where A is the asteroid's Bond albedo, S⊙ is the solar constant, q
is the angle between rays
from the Sun and the asteroid surface normal, s B is the Stefan-Boltzmann constant, h
is
the emissivity of the asteroid surface, mu is the atomic mass unit, L is the latent heat of
sublimation of the escaping gas, specified here as an energy per mole, and Z is the surface
sublimation rate (particles lost per unit area per unit time). For simplicity, some of these
quantities are fixed at constant values of A = 0.1, h = 1, and S⊙ = 1360 W m−2.
Although the resulting rate of mass loss is sensitive to details of the Sun–asteroid ge-
ometry, it is possible to discuss the behavior in two limiting cases for cosq (see Jewitt et al.
2011, 2015). A rapidly rotating asteroid will redistribute the incoming radiative energy from
its Sun-facing side over the majority of its emitting surface. In that case, a mean value of
cosq = 1/4 accounts for this efficient redistribution. On the other hand, a slowly rotating
(or non-rotating) asteroid will receive most of its energy at nearly sub-solar surface loca-
tions, and will emit sublimated gas only from the illuminated parts. Thus, one can assume
cosq = 1 at those locations. In these two cases the emitting surface area A of the asteroid
also differs. In the fast-rotating limit, nearly all points on the asteroid receive some heat, so
A ≈ p D2 (the full surface area). In the slow-rotating limit, the area is given roughly by the
front-side cross section, A ≈ p D2/4. This area is needed to compute the full mass loss rate
(9)
is the mean molar mass of the sublimating molecules. The quantity Zm mu is some-
where m
times called the mass erosivity e (mass lost per unit time per unit area).
M = Zm muA ,
Equation (8) has two undetermined parameters: T and Z. These can be found by using
the Clausius–Clapeyron relation, which specifies the vapor pressure Pv at the surface of a
sublimating body,
where P¥
is the vapor pressure in the high-temperature limit and kB is Boltzmann's con-
stant. The latent heat is assumed to be a constant, independent of temperature. Assuming a
Maxwellian distribution of escaping molecules, the mean sublimation rate at the surface is
given by the Hertz–Knudsen equation,
where g
e.g., van Lieshout et al. 2014).
is a dimensionless efficiency that takes account of a range of kinetic effects (see,
Table 2 lists many of the above properties for substances believed to exist on comets
and asteroids, listed in order of increasing latent heat. Figure 5(a) shows numerical solu-
tions for Z as a function of heliocentric distance ra, for a subset of the substances listed
in Table 2. In most cases there is an inverse monotonic relationship between L and Z at
any given distance. Close to the Sun, the sublimation term dominates the right-hand side
of Equation (8), and thus Z (cid:181)
r−2
a . At larger distances, Z drops off exponentially when the
radiative emission term begins to dominate. The fast-rotating limit was used to compute
cosq and A because observations (e.g., Campins et al. 2009) suggest inner heliospheric as-
teroids experience significantly more surface redistribution of thermal energy than predicted
by non-rotating models.
Figure 5(b) shows the associated energy-balance solutions for the surface temperature
T . For the most volatile substances, T is relatively flat near the Sun because the incoming
solar flux drives the sublimation phase change and does not heat up the asteroid. In that case,
Pv = P¥ exp(cid:18)−
muL
kBT(cid:19)
Z =
g Pv
p2pm mukBT
(10)
(11)
Predictions for Dusty Mass Loss from Asteroids
13
Table 2 Sublimation properties of cometary and asteroidal substances.
Substance
m
[g/mol]
r
[g/cm3]
L
[kJ/mol]
ln P¥
[ln Pa]
g
Refs
Carbon dioxide (CO2)
Water ice (H2O)
Sodium (Na)
Carbonaceous chondrite
Iron (Fe)
Silicon monoxide (SiO)
Fayalite (Fe2SiO4)
Enstatite (MgSiO3)
Forsterite (Mg2SiO4)
Quartz (SiO2)
Corundum (Al2O3)
Silicon carbide (SiC)
Graphite (C)
44.00
18.00
23.00
39.85
55.85
44.09
203.77
100.39
140.69
60.08
101.96
40.10
12.01
References: 1. Fanale and Salvail
Baldwin and Sheaffer (1974), 5. van Lieshout et al. (2014).
1.56
1.00
0.97
2.80
7.87
2.13
4.39
3.20
3.27
2.60
4.00
3.22
2.16
27.84
28.90
22.53
24.48
26.90
30.20
35.40
35.80
31.80
30.80
37.00
35.50
34.40
(1987), 2. Fanale and Salvail
26.34
48.06
105.80
320.67
402.03
411.72
501.99
572.92
542.99
577.38
643.24
652.36
778.60
1.0
1.0
1.0
1.0
1.0
0.04
0.1
0.1
0.1
1.0
0.1
0.1
0.1
1
2
3
4
5
5
5
5
5
5
5
5
5
(1984), 3. Huebner
(1970), 4.
the quantity 1/T varies approximately as (c1 + c2 lnra) for appropriate constants c1 and c2.
Farther from the Sun, however, T (cid:181)
r−1/2 in radiative equilibrium.
Because we do not yet know the detailed surface composition of active asteroids, it is
not clear how to specify the latent heat and other properties a priori. Sekanina (2003) mod-
eled the light curves of sungrazing comets by treating the latent heat L and mean molecular
mass m as constrained free parameters. The result of that process was a range of latent heats
(120 to 360 kJ mol−1) and molecular masses (200 to 800 g mol−1), with unique pairs of
values that successfully predict the light curve properties of each comet. Figure 6 compares
laboratory values of L and m
for the substances listed in Table 2 with those derived em-
pirically by Sekanina (2003). Shown for comparison are data from other listings of volatile
ices (Prialnik et al. 2004) and solid-phase elements (Huebner 1970). In this two-dimensional
plane, there is no real overlap between the laboratory values and those derived by Sekanina
(2003). However, it is important to note that the sublimation rate Z is driven mainly by the
value of L and only very weakly by m
m −1/2). Thus, the overlap that is seen in
just the L values may point to the existence of a heterogeneous mixture of substances with
a range of sizes and latent heats (e.g., hydrated silicates mixed with organic hydrocarbons)
that could be released from surfaces exposed to the near-Sun environment.
(i.e., Z (cid:181)
In a similar vein as the empirical study of Sekanina (2003), it is possible to use the
Jewitt et al. (2013) measurement of Phaethon's mass loss ( M ≈ 3 kg s−1 at perihelion) to
provide an observation-based estimate of L and m . However, the measured value of M from
Phaethon was a dust mass loss rate, whereas the quantity computed in Equation (9) cor-
responds to the gas/molecular component of the outflow. As stated above, there is still no
firmly accepted understanding of how active asteroids produce and eject dust grains. The gas
and dust components may be intimately connected to one another (via, e.g., re-condensation
of sublimated molecules) or they may originate from completely different regions on the
surface (see Jewitt et al. 2015). However, we can note that many comets are inferred to have
dust-to-gas mass ratios M centered around unity (e.g., A'Hearn et al. 1995; Sanzovo et al.
1996; Kolokolova et al. 2007). Thus, we make a trial assumption of M = 1, which allows
us to take M to be equal to the dust mass loss rate. The assumed value of M is likely to
be an important source of uncertainty for the model results given below. The sensitivity of
these results to changes in M is explored further below as well.
14
S. R. Cranmer
Fig. 5 Dependence on asteroid heliocentric distance ra of: (a) surface sublimation rates Z and (b) energy-
balance temperatures T , plotted for a subset of the substances listed in Table 2. Unlabeled curves correspond
to atomic iron (orange dashed curve), crystalline enstatite (red dotted curve), and corundum (red solid curve).
See text for discussion of the empirical model (black solid curve).
An empirical model of the sublimation of material with arbitrary L and m
requires spec-
ification of the other sublimation properties of the material. We follow Huebner (1970) and
Sekanina (2003) by defining
P¥ = P0 exp(cid:18) muL
kBT0(cid:19)
(12)
with a fiducial value of P0 = 0.398 GPa. The result of computing T0 for the 13 actual sub-
stances listed in Table 2 is a mean value (after excluding CO2 and H2O) of 5300 K, which
is used in the models below.3 Since the assumed heterogeneous mixture of substances may
contain both dusty ices (g ≈ 1) and silicates (g ≈ 0.1) we assume a mean value of g = 0.3.
3 This value is also consistent with the Clausius–Clapeyron relation used by Sekanina (2003).
Predictions for Dusty Mass Loss from Asteroids
15
Fig. 6 Comparison of laboratory measurements of the sublimation properties of solid substances (red filled
circles from Table 2; gold crosses from Huebner 1970; blue triangles from Prialnik et al. 2004) with empirical
determinations from sungrazing comet light curves (green diamonds from Tables 1 and 3 of Sekanina 2003).
Contours that reproduce the measured mass loss of Phaethon are shown in black (solid: fast-rotating limit,
dashed: slow-rotating limit).
A comprehensive search of the two-dimensional space shown in Figure 6 was under-
taken for a Phaethon-like asteroid at perihelion, with ra = 0.14004 AU and D = 5.12 km.
Values of L and m
that yielded M = 3 kg s−1 are shown by the black curves in Figure 6;
one computed for the fast-rotating limit and one for the slow-rotating limit. For simplic-
ity, we choose one point along the fast-rotating locus of solutions for use in the models
presented below: L = 204 kJ mol−1 and m = 100 g mol−1 (see the black filled square in
Figure 6). This value of L falls comfortably within the empirically determined range found
by Sekanina (2003) for sungrazing comets. Also, our solutions are close to the value of 320
kJ mol−1 listed in Table 2 for carbonaceous chondrites. It should be noted that the value of L
for chondritic material is not really known so precisely. Both Baldwin and Sheaffer (1974)
and Chyba et al. (1993) reported approximate values for the heat of ablation of chondrite
meteors spanning values between 200 and 350 kJ mol−1. To repeat, we believe this empiri-
cal solution may point to the existence of a mixture of multiple solid species that, taken in
bulk, sublimate at a similar rate as a single compound with representative values of L and m .
Figure 7 shows computed mass loss rates versus ra for this paper's collection of 97
closest-approach events with SPP, and we use the empirical solution described above: L =
204 kJ mol−1 and m = 100 g mol−1 in the fast-rotating limit. When these parameters are
held fixed, it is clear that ra is the parameter most important for determining M, with the
asteroid diameter D providing a much more limited variation. The red filled circle indicates
the position of asteroid Phaethon at its closest approach with SPP, and the red X symbol
shows Phaethon at its perihelion. At the time of closest approach with SPP, Phaethon's mass
16
S. R. Cranmer
Fig. 7 Gas mass loss rates M for asteroids at their times of closest approach with SPP (symbols, with colors
corresponding to H with same scaling as in Figure 2) and for asteroid 3200 Phaethon at various points along
its orbit (red solid curve) and at perihelion (red X symbol). A similar model for comet C/2003 K7 (brown
dotted curve) is compared to the measured range of dust M values from Ciaravella et al. (2010) (brown strut).
Non-detection upper limits from Keck observations of active asteroids far from perihelion (Jewitt 2013) are
shown with black arrows.
loss rate is predicted to be almost a million times lower than when it reaches perihelion.
This again provides a warning that the times of closest approach may not be the most aus-
picious observation times. Observations of Phaethon and several other asteroids taken at
larger distances (ra = 1.4–2.6 AU; see Jewitt 2013) showed no discernible mass loss. The
approximate upper limits on M shown in Figure 7 are many orders of magnitude larger than
what our model would predict at those distances.
Another comparison between observed dust mass loss and the model described above is
shown in Figure 7. Ciaravella et al. (2010) observed silicon and carbon ultraviolet emission
from sungrazing comet C/2003 K7 when it was at a distance of ra = 0.016 AU. They de-
duced a range of dust mass loss rates assuming the silicon comes from the dissociation of
molecules like olivine or forsterite. Ciaravella et al. (2010) also inferred a diameter for the
nucleus of about 0.06–0.12 km. Using a central value of D = 0.1 km and the fast-rotating
M for the gas that fall within the
model above with L = 204 kJ mol−1, we found values of
uncertainty limits of the Ciaravella et al. (2010) dust mass loss rate.
The agreement between the above model and the Ciaravella et al. (2010) observation
represents additional support for M ≈ 1 for inner heliospheric mixtures of dust and gas in
the vicinity of sublimating bodies. However, if we had chosen other values of M , it would
have only required a small change in the latent heat to reproduce the Jewitt et al. (2013)
mass loss rate. Specifically, the standard value of L = 204 kJ mol−1 (which was optimized
Predictions for Dusty Mass Loss from Asteroids
17
for M = 1) needs to be decreased only to 172 kJ mol−1 for M = 0.01, or increased to 237
kJ mol−1 for M = 100. These changes in L have been incorporated into an approximate
fitting formula for the sublimation rate Z as a function of both heliocentric distance and the
dust-to-gas mass ratio:
Zfit(ra, M ) = Z0 M −0.0427 r−2.25
a
,
(13)
where ra is expressed in AU, Z0 = 2.545 ×1016 particles cm−2 s−1, and s = 0.0787M −0.0661.
This fitting formula is used below in order to vary M as a free parameter while retaining
the empirical calibration to Phaethon's dust mass loss rate.
(cid:2)1 + (ra/s )3(cid:3)−6
5 Coma and Tail Formation
The goal of this section is to estimate a representative length scale for the dust-filled coma
or tail surrounding an active asteroid in the inner heliosphere. This length scale is defined
specifically as the largest distance from the asteroid at which WISPR on SPP is expected
to see a dust-scattered enhancement over the sky background. As above, we continue to use
the observations of Phaethon at its perihelion (Jewitt et al. 2013) to help constrain some of
the unknown parameters of the model.
5.1 Spherically Symmetric Model
Huebner (1970) and Delsemme and Miller (1971) computed the maximal radial distance
traversed by dust grains that are ejected from a comet's surface. Newly freed grains become
exposed immediately to sunlight and begin to sublimate in the same manner as the parent
asteroid. A spherical dust grain with radius a should have a finite lifetime t given by
t =
r a
m muZ
,
(14)
where r
is the mean mass density of escaping material. In the models below, we assume
a representative value r = 3 g cm−3. A dust grain that drifts away from the parent body
with velocity vdust will thus traverse a radial distance of order R ≈ vdustt before it sublimates
away completely. However, Jewitt et al. (2013) speculated that the observed radial distance
of coma-like emission is probably going to be smaller than R, because instrumental effects
(i.e., a high sky background and low photon counting statistics) can obscure the faint outer
parts of the dust cloud. Thus, we need to estimate the visible-light brightness of a dust coma
(as a function of impact-parameter distance b away from the asteroid) and compare it to the
sky background flux FZ defined in Section 3.
In lieu of a full three-dimensional model of the dynamics of dust grains leaving the
asteroid, an approximate spatial distribution can be computed using the model of Haser
(1957). Making use of mass flux conservation in spherical symmetry, with a sink term to
account for grain loss via sublimation, results in a time-steady solution for the dust number
density,
r (cid:19)2
nd(r) = nd0(cid:18) Ra
exp(cid:20)−(cid:18) r − Ra
vdustt (cid:19)(cid:21) ,
(15)
18
S. R. Cranmer
where Ra is the asteroid radius and r is the asteroid-centric radial distance at which nd is
measured. The value of nd0 at the asteroid surface can be estimated if we know the dust-to-
gas mass ratio M and the number density of gas molecules ng0 at the surface,
nd0 =
(m mu) ng0
md
M
(16)
and md, the mass of an individual dust grain, can be computed straightforwardly given
its mean density r , radius a, and the assumption it is roughly spherical in shape. The
Maxwellian gas mass loss theory (Equation 11) also specifies
where the mean speed of gas molecules is given by
ng0 = 4Z / ¯v
¯v = s 8kBT
pm mu
(17)
(18)
(e.g., Delsemme and Miller 1971).
If a telescope views the dust emission at an impact-parameter distance b away from the
asteroid itself, it will see dust grains with a given column density N(b). To compute the
column density, we assume the grains are distributed with enough empty space around them
so their observed cross sections do not overlap. Thus, N(b) is given by an optically thin
integral over the line-of-sight distance x,
N(b) = Z +¥
−¥
dx nd(x) =
nd0 R2
a
b
ILOS(b )
where the dimensionless line-of-sight (LOS) integral is defined as
(19)
(20)
ILOS(b ) = Z +¥
−¥
du
1 + u2 exp(cid:16)−b p1 + u2(cid:17)
and b = b/(vdustt ). The above integral was derived by defining the heliocentric distance r of
each point along the LOS using r2 = x2 + b2, and the dimensionless integration coordinate
is u = x/b. The above expression also assumes that b ≫ Ra. The integral was computed
numerically over a fine grid in the parameter b , and this was used as a lookup table when
computing actual column densities. Figure 8 shows how ILOS varies with b .
The visible-light flux Fdust emitted by dust grains (at a given distance b away from the
asteroid, and in a solid angle W
that fills a WISPR pixel) can be computed and compared
to the background sky flux. The total number of grains that contribute to this quantity is the
product of the column density N(b) and the cross-sectional area of one pixel as viewed by
the spacecraft. If the spacecraft–asteroid distance d is assumed to be much larger than the
impact parameter b, the relevant cross-sectional area can be estimated to be W d2. Thus, if a
single dust grain emits a visible-light flux F1, the total flux in the pixel is given by
Fdust = F1 N(b)W d2 .
(21)
The calculation of F1 was done using the equations in Section 3 and the assumption that
each dust grain is essentially a "tiny asteroid" with a comparable albedo to its parent body.
At any time during an encounter between an asteroid and SPP, we can compute Fdust
over a range of trial values of the impact parameter b. It is then possible to solve for a
Predictions for Dusty Mass Loss from Asteroids
19
Fig. 8 Numerically integrated values of ILOS versus b (solid black curve). The dominant exponential behavior
is removed by multiplying ILOS by the quantity eb p1 + b
(red dashed curve).
representative observed coma extent (bcoma) that is the distance at which Fdust is equal to a
threshold sky background flux. That sky flux was given by 10−4FZ, which was determined
in Section 3 to be the practical limit in modern-day heliospheric imagers for resolving small
features from a large-scale background (DeForest et al. 2011). The quantity FZ depends on
the elongation angle e and the SPP heliocentric distance rp. Thus, the derived value of bcoma
is observer-dependent and not intrinsic to the asteroid.
In the limiting case of b ≪ 1, the dimensionless integral ILOS approaches a constant
value of p . Thus, the solution for the observed coma size can be written as
bcoma ≈ (cid:18) F1
10−4 FZ(cid:19) p R2
a nd0 W d2 .
(22)
The above expression was used as a validation for the full numerical solution in the limit of
bcoma ≪ vdustt .
The model described above has three parameters that still have not yet been specified:
M , vdust, and a. We will use the Jewitt et al. (2013) measurement of bcoma = 2.5 × 105 km
for Phaethon at its perihelion in order to constrain them. Plausible ranges of variability for
these parameters are given as follows:
1. The dust-to-gas mass ratio M has been discussed above. Comet observations appear
to give values of M ∼ 1 (A'Hearn et al. 1995; Sanzovo et al. 1996; Kolokolova et al.
2007), and recent high-quality measurements of comet 67P/Churyumov-Gerasimenko
indicate values between 3 and 10 (Rotundi et al. 2015; Fulle et al. 2016). Measurements
for active asteroids do not yet exist. Thus, with no firm firm observational guidance, we
will vary M widely over seven orders of magnitude, between values of 10−3 and 104.
20
S. R. Cranmer
2. The grain velocity vdust can be parameterized as a ratio x = vdust/vgas, since vgas is
known from the sublimation model. Delsemme and Miller (1971) found that escaping
gas molecules eventually accelerate to an asymptotic speed of vgas ≈ 1.8 ¯v. Thus, we
specify vdust as a function of ¯v and x . For asteroids in the inner heliosphere, the tem-
peratures shown in Figure 5(b) indicate vgas ≈ 0.3–1 km s−1. Comet observations tend
to show that vgas is a practical upper limit for vdust (e.g., Waniak 1992; Hughes 2000;
Prialnik et al. 2004; Beer et al. 2006; Bonev et al. 2008; Jewitt 2012; Ishiguro et al. 2016).
However, vdust can take on a range of smaller values as well. A lower limit on vdust is
the escape velocity vesc = (2GMa/Ra)1/2, which is typically between 10−4 and 10−3
km s−1 for the asteroids considered here. Thus, we will assume a plausible range for the
dimensionless velocity ratio of 10−4 < x < 1.
3. Astrophysical dust tends to exhibit a broad distribution of grain radii a rather than any
one specific value (Combi 1994; Fulle 2004). Observations of comets and the zodiacal
light indicate particle sizes spanning the range from 0.1 m m to 10 cm (Fulle et al. 1993;
Mann et al. 2004; Horz et al. 2006; Kretke and Levison 2015). Some models of dust loss
from comets show a distinct anticorrelation between a and vdust (e.g., Finson and Probstein
1968; Delsemme and Miller 1971; Wallis 1982), and this can be used to estimate the
value of a that maximizes the drift distance R ≈ vdustt . However, observations often
show grain sizes in excess of this putative maximum value (Tenishev et al. 2011). These
models also underestimate vdust by giving values only slightly larger than vesc, whereas
observations often show a range of higher speeds extending up to vgas (see above). In
order to avoid undue dependence on these kinds of models, we allow the value of a to
vary freely between 0.01 m m and 10 cm.
With the above parameters (M , vdust, a) varied freely and the other asteroid properties
fixed for Phaethon at its perihelion, it is straightforward to compute bcoma using Equations
(14)–(21) and compare the results to the observed value of 2.5 × 105 km. An initial search
of the parameter space yielded several clear constraints on the parameters. To match the ob-
served coma size, models with x ≤ 1 must have a dust-to-mass ratio of M > 0.48 and a rep-
resentative grain size of a < 39 m m. This latter constraint is in agreement with Jewitt et al.
(2013), who inferred a grain radius of a ≈ 1 m m for the particles producing the emission in
the observed coma of Phaethon.
In order to help better constrain valid ranges of the free parameters, we searched for
degeneracies (i.e., combinations of parameters that produced identical values of bcoma). Fig-
ure 9 shows there is an extremely narrow "allowed" region in a two-dimensional cut through
the solution space when the orthogonal axis parameters are defined as
C1 = x M , C2 = ax −0.15 .
(23)
The points shown in Figure 9 are a subset of results from a Monte Carlo simulation of 106
random trial solutions. The 5363 displayed points represent only those solutions that agree
with the observed value of bcoma to within 1%.
Because all of the points shown in Figure 9 reproduce the SECCHI observations of
bcoma for Phaethon at perihelion, we proceed by choosing one arbitrary set of values from
those points to use in the models below. We presume the results for other asteroids at other
heliocentric distances should scale similarly no matter the details of this choice. We follow
Jewitt et al. (2013) by adopting a = 1 m m as a mean dust grain radius. With that constraint,
the other two parameters are seen to follow the relationship x ≈ 1.76/M 1.085. Optimized
values of M = 3 and x = 0.53 were thus chosen in order to maintain continuity with the
cometary analogy; i.e., the empirical knowledge that M tends to be between 1 and 10 and
that vdust ≈ vgas (or at least vdust ≫ vesc) is often seen (e.g., Jewitt 2012; Ishiguro et al. 2016).
Predictions for Dusty Mass Loss from Asteroids
21
Fig. 9 Two-dimensional parameter space of solutions to for bcoma that agree with observations of Phaethon
at its perihelion. Symbol colors correspond to the values of x (see legend), and the gray cross indicates values
adopted for use with the SPP asteroid data: M = 3, x = 0.53, a = 1 m m.
5.2 Results for SPP Asteroid Encounters
Figure 10 shows some representative calculations of bcoma for the set of 97 closest-approach
events described above (filled circles), and for a set of idealized asteroid properties and
positions (solid curves). For these idealized curves, it was assumed that the three relevant
bodies (the asteroid, SPP, and the Sun) were situated on the corners of an equilateral triangle.
In other words, for each point along these curves, ra = rp = d and a = e = 60◦. For the 97
closest-approach events, these geometrical properties were extracted from the ephemerides
discussed earlier. In cases when the dust ejected by an asteroid does not ever emit enough
flux to exceed the observable sky background, we set bcoma to a lower limit of the asteroid
radius Ra. This happens for all modeled cases at ra > 0.3 AU, which helps to justify our
choice for disregarding asteroids with perihelia q > 0.3 AU.
Note that there is a relatively finite range of heliocentric distances inside of which an
asteroid is expected to emit grains that survive for thousands of kilometers and produce a
bright dust coma. This range appears to be roughly 0.08 < ra < 0.25 AU. For asteroids closer
to the Sun than about 0.05 AU, the sublimation rate Z is so high that the grain lifetime t
is
extremely short. Thus, the grains cannot reach large values of b before they are destroyed.
For asteroids further away from the Sun than about 0.3 AU, the sublimation rate Z drops off
to exceedingly small values. This allows any escaping grains to essentially "live forever,"
but their number density nd is too low for their flux to compete effectively with the sky
background.
Rather than limit the calculation to the small database of 97 closest-approach events, the
entire ephemeris for each pairing of SPP with a given asteroid was processed to compute
22
S. R. Cranmer
Fig. 10 Distance dependence of the modeled observable dust cloud size bcoma. Symbols correspond to the 97
closest-approach events, with colors corresponding to H (i.e., asteroid diameter D) with the same scalings as
in Figure 2. Curves correspond to grids of models with idealized geometrical configurations and fixed asteroid
diameters (see text). The measured tail length of Phaethon at its perihelion (Jewitt et al. 2013) is shown with
a red X symbol.
bcoma as a detailed function of time (i.e., at 0.1 day intervals). To compare directly with
planned observations with WISPR, each value of bcoma was converted into a sky angle q coma
as observed from the vantage point of SPP,
q coma = tan−1 (bcoma/d) .
(24)
Because bcoma is a radius and not a diameter, a spherical dust cloud should have an observ-
able angular extent of 2q coma. However, we provide q coma as a conservative lower limit in
cases of efficient tail "blowback" behind the asteroid. A dust cloud is thus considered to be
resolvable only when q coma exceeds the size of a single WISPR pixel (q ≈ 1.45′).
Figure 11 shows the full set of cases with modeled values of q coma that exceed 0.5′ on
the sky. The local maxima in q coma occur neither at the times when d = dmin nor at the times
when ra = q. The large asteroid Phaethon shows up prominently with repeated peak values
of q coma between 10′ and 26′, no matter the location of SPP in the inner heliosphere. The
smaller asteroids 137924, 155140, and 289227 each have one-time favorable events with
q coma > 17′, but at other perihelion passes the angular extents are smaller. Figure 11 also
makes clear that dust tail sizes greater than about 1′ only tend to occur when asteroids fall
between about 0.1 and 0.2 AU. Together with the results shown in Figure 10, this supports
our choice to disregard asteroids with perihelia greater than 0.2–0.3 AU.
Figure 12 shows essentially the same information that is in Figure 11, but plotted as a
function of mission time t instead of heliocentric distance. This shows that the individual
Predictions for Dusty Mass Loss from Asteroids
23
Fig. 11 Evolution of the modeled dust-cloud angular size q coma versus asteroid heliocentric distance for the
full set of modeled encounters with SPP. Curve colors correspond to H (i.e., asteroid diameter D) with the
same scalings as earlier figures. The fiducial WISPR pixel size of 1.45′ is noted with a dotted line.
episodes of large observable angular extent are limited in time and distributed sporadically
through the SPP mission. There are 113 predicted maxima that exceed the assumed WISPR
pixel size, corresponding to 24 distinct asteroids. Several of the maxima are closely spaced
pairs or triplets, separated by hours to days, but most are isolated in time. The mean duration
of an event (defined as the time spent with q coma ≥ 1.45′) is 1.39 days, but the events cover
a range from 0.2 to 5.8 days. Table 3 lists the most promising 41 of these events (i.e., only
those with q coma ≥ 3′) in order of mission time, and also gives the apparent magnitude mV
of the parent asteroid and its heliocentric distance at the specific times of maximum q coma.
6 Discussion and Conclusions
The goal of this paper is to call the community's attention to the likelihood that SPP will
be well-positioned to observe mass loss from Mercury-crossing asteroids in the inner helio-
sphere. Specifically, we predict that there will be several times during the SPP mission when
the WISPR instrument will be able to detect visible-light emission from the asteroids them-
selves and (in a few cases) from associated dust clouds that may subtend almost a degree of
angular width on the sky. These observations could fill in a large gap between the properties
of two heretofore distinct populations-active asteroids and sungrazing comets-and thus
help complete the census of primordial solar system material.
Because of several ongoing uncertainties, many of the quantitative predictions made
above may not remain valid for the actual SPP mission to commence in 2018. Specifically,
the following four factors will need to be re-evaluated over the coming few years:
24
S. R. Cranmer
Fig. 12 Evolution of the modeled dust-cloud angular size q coma for various asteroids, versus mission time t
in days. Curve colors correspond to H (i.e., asteroid diameter D) with the same scalings as earlier figures.
The fiducial WISPR pixel size of 1.45′ is noted with a dotted line.
1. As noted in Section 2, if the spacecraft launch slips from its nominal date of July 31,
2018, the predicted time-dependent distances between SPP and the asteroids will be in-
correct. The mission depends on multiple close encounters with Venus to adjust the tra-
jectory into the desired elliptical orbit with a perihelion of 0.0459 AU. Fox et al. (2015)
described a backup plan that involves a May 2019 launch plus one additional Venus
gravity assist to bring SPP into an orbit similar to the baseline trajectory.
2. Many of the objects in our database of 97 Mercury-crossing asteroids have been dis-
covered relatively recently. Thus, there may still be substantial uncertainties in their
ephemeris parameters. Such errors would necessarily propagate into our predictions of
the relative times and distances of encounters with SPP. Whether the improvement of
these parameters would give rise to a larger or smaller number of favorable encounters
remains to be seen. Nevertheless, the pace of asteroid discovery is likely to continue,
and there may be dozens more possible targets discovered between now and 2018.
3. The predictions made above did not take into account that the WISPR instrument has a
finite field of view and cannot see the entire sky. Some fraction of favorable encounters
with asteroids may end up being hidden behind the SPP heat shield or other parts of
the spacecraft. Thus, not every asteroid in listed Tables 1 and 3 will be observable at
all times. These details need to be considered when constructing detailed observation
plans, but they are beyond the scope of this paper.
4. The observability of any given asteroid and its surrounding dust cloud was computed
using a threshold sky background of 10−4FZ (see, e.g., DeForest et al. 2011). It is possi-
ble that WISPR may contain sufficient improvements in photon counting or flat-fielding,
relative to the SECCHI package on STEREO, that could allow even weaker signals to be
extracted from the raw images.
However the above issues are resolved, there appears to be a high probability for a significant
number of encounters between SPP and Mercury-crossing asteroids during times when the
Predictions for Dusty Mass Loss from Asteroids
25
Table 3 Predictions for large angular dust tail events, sorted by SPP mission time.
Time [days]
Name
187.90
230.71
234.80
246.21
293.41
336.51
339.61
593.10
770.76
790.45
857.14
860.34
863.24
893.13
897.13
897.23
899.03
899.83
994.40
994.60
1192.76
1210.85
1291.13
1384.41
1385.01
1386.91
1492.68
1493.88
1496.98
1590.26
1682.44
1709.33
1792.21
1796.71
1907.58
2091.79
2328.12
2377.97
2391.48
2416.71
2431.22
394130
394392
394392
374158
137924
Phaethon
Phaethon
137924
374158
394392
2008 HW1
Phaethon
Phaethon
137924
289227
137924
289227
289227
2008 MG1
2008 MG1
137924
2006 TC
374158
Phaethon
Phaethon
Phaethon
137924
2005 HC4
137924
155140
2013 HK11
394130
137924
137924
Phaethon
137924
374158
2008 HW1
137924
431760
Phaethon
q coma [arcmin]
3.642
4.089
3.044
6.120
3.456
10.813
7.634
3.851
3.226
3.515
3.880
12.624
10.725
4.837
19.625
3.427
6.825
7.432
3.569
3.549
5.927
4.175
3.787
15.789
14.864
26.335
17.238
5.596
9.264
22.926
3.211
3.302
5.994
5.737
15.775
4.344
4.232
3.645
4.637
6.875
24.707
mV
14.8
14.8
15.1
13.6
15.6
12.1
13.8
15.2
15.2
14.9
14.9
11.8
13.1
11.1
11.2
10.4
12.8
13.4
14.3
14.2
13.1
14.7
14.8
10.5
10.6
10.9
9.6
13.5
12.2
9.0
12.5
13.6
14.2
14.3
12.1
15.0
13.7
14.4
14.7
13.1
11.1
ra [AU]
0.14501
0.13994
0.13974
0.14030
0.14530
0.15449
0.14855
0.14837
0.14133
0.13631
0.14726
0.15416
0.15071
0.13828
0.13989
0.13533
0.12748
0.13680
0.14497
0.14520
0.14399
0.14249
0.14196
0.15016
0.14506
0.15667
0.13711
0.13366
0.14556
0.16850
0.15668
0.14533
0.14426
0.14657
0.15692
0.14860
0.14082
0.14451
0.14859
0.14104
0.15633
latter may be losing mass at ra < 0.2 AU. Details aside, the statistical distribution of events
is likely to remain similar to what was computed in this paper.
In addition to refining the positional and temporal accuracy of the above predictions,
there are also several ways that the mass loss modeling can be improved. Our assumption of
constant values for L and m should be replaced by a more self-consistent (i.e., temperature
dependent) description of specific ejected materials. This is particularly important because
the computed rates are often in the exponentially dropping part of the sublimation curve
(Figure 5), where small variations in the input parameters could change Z by several orders
of magnitude. Also, the spherically symmetric Haser (1957) model should be replaced by
a full three-dimensional dynamical simulation of the ejected dust grains. If the escaping
26
S. R. Cranmer
grains are swept back into a collimated tail, their number density may be up to an order of
magnitude higher (and thus more easily observable when viewed from a favorable direction)
than if they were spread out in a spherical cloud. Jewitt et al. (2011) estimated the sunward
turnaround distance s expected from grains of a given size. In cases where s ≪ bcoma it would
be most useful to apply such a three-dimensional correction to the density model. At the very
least, the standard Finson and Probstein (1968) type of ballistic modeling should be done for
the most promising encounters, in order to predict the tail orientations and geometries.
The idea of using inner heliospheric space probes as remote observatories for active
asteroids should be expanded beyond just SPP. The Solar Orbiter mission (Muller et al.
2013; Bemporad et al. 2015) will reach a minimum perihelion distance of about 0.28 AU
during a similar time-frame as SPP, but it will also leave the ecliptic plane to eventually reach
inclination angles of order 30◦. Sarli et al. (2015) proposed a mission to visit Phaethon and
associated asteroids (155140) 2005 UD and (225416) 1999 YC, which would explore the
origins of the Geminid meteor stream and study the physics of comet/asteroid transition
objects. Near-Earth asteroid flyby or rendezvous missions with infrared spectro-imagers
(Groussin et al. 2016) could also improve our knowledge of the physics of regolith loss in a
hot thermal environments. Lastly, any spacecraft that comes close enough to fly through the
dust tail of an active asteroid would put unprecedented constraints on the properties of the
ejected grains, thus allowing a comparison of similarities and differences to dust ejected by
comets (see, e.g., Kissel and Krueger 1987; Mann and Czechowski 2005; Neugebauer et al.
2007; Della Corte et al. 2015).
Acknowledgements The author gratefully acknowledges Kelly Korreck and Martha Kusterer for supplying
the SPP SPICE kernel. The initial idea for this paper grew out of online discussions with Corey Powell
and Matthew Francis about interesting new destinations for space probes after the July 2015 New Horizons
flyby of Pluto. The author also thanks David Malaspina for helpful comments on the manuscript, and the
anonymous referees for many constructive suggestions. This work was supported by start-up funds from the
Department of Astrophysical and Planetary Sciences at the University of Colorado Boulder. This research
made extensive use of NASA's Astrophysics Data System.
References
C.H. Acton, Plan. Space Sci. 44, 65 (1996)
J. Agarwal, D. Jewitt, H. Weaver, et al., Astron. J. 151, 12 (2016)
M.F. A'Hearn, R.L. Millis, D.G. Schleicher, et al., Icarus 118, 223 (1995)
C.W. Allen, Astrophysical Quantities, 3rd ed. (London: Athlone Press, 1973)
B. Baldwin, Y. Sheaffer, J. Geophys. Res. 76, 4653 (1974)
E.H. Beer, M. Podolak, D. Prialnik, Icarus 180, 473 (2006)
A. Bemporad, S. Giordano, J.C. Raymond, et al., Adv. Space Res. 56, 2288 (2015)
D.A. Biesecker, P. Lamy, O.C. St. Cyr, et al., Icarus 157, 323 (2002)
T. Bonev, K. Jockers, N. Karpov, Icarus 197, 183 (2008)
J. Borovicka, Z. Charv´at, Astron. Astrophys. 507, 1015 (2009)
E. Bowell, B. Hapke, D. Domingue, et al., in Asteroids II, ed. R. Binzel, T. Gehrels, M. Matthews (Tucson:
University of Arizona Press), p. 524 (1989)
J.C. Brandt, M. Snow, Icarus 148, 52 (2000)
P. Bryans, W.D. Pesnell, Astrophys. J. 760, 18 (2012)
H. Campins, M.S. Kelley, Y. Fern´andez, et al., Earth Moon Planets 105, 159 (2009)
C.F. Chyba, P.J. Thomas, K.J. Zahnle, Nature 361, 40 (1993)
A. Ciaravella, J.C. Raymond, S. Giordano, Astrophys. J. 713, L69 (2010)
M.R. Combi, Astron. J. 108, 304 (1994)
C.E. DeForest, T.A. Howard, S.J. Tappin, Astrophys. J. 738, 103 (2011)
M. Delbo, G. Libourel, J. Wilkerson, et al., Nature 508, 233 (2014)
V. Della Corte, A. Rotundi, M. Fulle, et al., Astron. Astrophys. 583, A13 (2015)
Predictions for Dusty Mass Loss from Asteroids
27
A.H. Delsemme, D.C. Miller, Plan. Space Sci. 19, 1229 (1971)
C.J. Eyles, R.A. Harrison, C.J. Davis, et al., Solar Phys. 254, 387 (2009)
F.P. Fanale, J.R. Salvail, Icarus 60, 476 (1984)
F.P. Fanale, J.R. Salvail, Icarus 72, 535 (1987)
M.J. Finson, R.F. Probstein, Astrophys. J. 154, 327 (1968)
N.J. Fox, M.C. Velli, S.D. Bale, et al., Space Sci. Rev. online first, DOI 10.1007/s11214-015-0211-6 (2015)
M. Fulle, in Comets II, ed. M. Festou, H. Keller, H. Weaver (Tucson: University of Arizona Press), p. 565
(2004)
M. Fulle, V. Mennella, A. Rotundi, et al., Astron. Astrophys. 276, 582 (1993)
M. Fulle, F. Marzari, V. Della Corte, et al., Astrophys. J. 821, 19 (2016)
J.D. Giorgini, in New Challenges for Reference Systems and Numerical Standards in Astronomy, ed. N. Cap-
itaine (Obs. de Paris), p. 87 (2011)
J.D. Giorgini, D.K. Yeomans, A.B. Chamberlin, et al., Bull. Am. Astron. Soc. 28 (3), 1158 (1996)
O. Groussin, J. Licandro, J. Helbert, et al., Exp. Astron. 41, 95 (2016)
B. Gundlach, J. Blum, Astron. Astrophys. 589, A111 (2016)
W.K. Hartmann, D.J. Tholen, K.J. Meech, et al., Icarus 83, 1 (1990)
L. Haser, Bull. Acad. Roy. Belgique 43, 740 (1957)
F. Horz, R. Bastien, J. Borg, et al., Science 314, 1716 (2006)
R.A. Howard, J.D. Moses, A. Vourlidas, et al., Space Sci. Rev. 136, 67 (2008)
H.H. Hsieh, D.C. Jewitt, Y.R. Fern´andez, Astron. J. 127, 2997 (2004)
W.F. Huebner, Astron. Astrophys. 5, 286 (1970)
W.F. Huebner, D.C. Boice, N.A. Schwadron, Adv. Space Res. 39, 413 (2007)
D.W. Hughes, Plan. Space Sci. 48, 1 (2000)
M. Ishiguro, Y. Sarugaku, D. Kuroda, et al., Astrophys. J. 817, 77 (2016)
P. Jenniskens, Earth Moon Planets 102, 505 (2008)
D. Jewitt, Astron. J. 143, 66 (2012)
D. Jewitt, Astron. J. 145, 133 (2013)
D. Jewitt, H. Hsieh, J. Agarwal, in Asteroids IV, ed. P. Michel, F. DeMeo, W. Bottke (Tucson: University of
Arizona Press), in press, arXiv:1502.02361 (2015)
D. Jewitt, J. Li, J. Agarwal, Astrophys. J. 771, L36 (2013)
D. Jewitt, H. Weaver, M. Mutchler, et al., Astrophys. J. 733, L4 (2011)
T. Kasuga, T. Yamamoto, H. Kimura, et al., Astron. Astrophys. 453, L17 (2006)
M.S. Kelley, D.J. Lindler, D. Bodewits, et al., Icarus 222, 634 (2013)
M.S. Kelley, D.J. Lindler, D. Bodewits, et al., Icarus 262, 187 (2015)
J. Kissel, F.R. Krueger, Nature 326, 755 (1987)
L. Kolokolova, H. Kimura, N. Kiselev, et al., Astron. Astrophys. 463, 1189 (2007)
K.A. Kretke, H.F. Levison, Icarus 262, 9 (2015)
S.M. Kwon, S.S. Hong, J.L. Weinberg, New Astron. 10, 91 (2004)
C.-I. Lagerkvist, P. Magnusson, Astron. Astrophys. Suppl. 86, 119 (1990)
C. Leinert, Space Sci. Rev. 18, 281 (1975)
A. Malhotra, J. D. Mathews, J. Geophys. Res. 116, A04316 (2011)
I. Mann, A. Czechowski, Astrophys. J. 621, L73 (2005)
I. Mann, H. Kimura, D.A. Biesecker, et al., Space Sci. Rev. 110, 269 (2004)
B.G. Marsden, Ann. Rev. Astron. Astrophys. 43, 75 (2005)
T. Matsakos, A. Uribe, A. Konigl, Astron. Astrophys. 578, A6 (2015)
E. Mazzotta Epifani, M. Dall'Ora, D. Perna, et al., Mon. Not. Roy. Astron. Soc. 415, 3097 (2011)
D.J. McComas, M. Velli, W.S. Lewis, et al., Rev. Geophys. 45, RG1004 (2007)
A. Morbidelli, R. Jedicke, W.F. Bottke, et al., Icarus 158, 329 (2002)
K. Muinonen, E. Bowell, K. Lumme, Astron. Astrophys. 293, 948 (1995)
D. Muller, R.G. Marsden, O.C. St. Cyr, H.R. Gilbert, Solar Phys. 285, 25 (2013)
R.H. Munro, B.V. Jackson, Astrophys. J. 213, 874 (1977)
A. Mura, P. Wurz, J. Schneider, et al., Icarus 211, 1 (2011)
M. Neugebauer, G. Gloeckler, J.T. Gosling, et al., Astrophys. J. 667, 1262 (2007)
M.S. Povich, J.C. Raymond, G.H. Jones, et al., Science 302, 1949 (2003)
G.W. Preston, Astrophys. J. 147, 718 (1967)
D. Prialnik, J. Benkhoff, M. Podolak, in Comets II, ed. M. Festou, H. Keller, H. Weaver (Tucson: University
of Arizona Press), p. 359 (2004)
S. Protopapa, J.M. Sunshine, L.M. Feaga, et al., Icarus 238, 11 (2014)
S. Rappaport, A. Levine, E. Chiang, et al., Astrophys. J. 752, 1 (2012)
J.C. Raymond, P.I. McCauley, S.R. Cranmer, et al., Astrophys. J. 788, 152 (2014)
28
S. R. Cranmer
A. Rotundi, H. Sierks, V. Della Corte, et al., Science 347, aaa3905 (2015)
R. Sanchis-Ojeda, S. Rappaport, E. Pall`e, et al., Astrophys. J. 812, 112 (2015)
G.C. Sanzovo, P.D. Singh, W.F. Huebner, Astron. Astrophys. Suppl. 120, 301 (1996)
B.V. Sarli, Y. Kawakatsu, T. Arai, J. Spacecraft Rockets 52, 739 (2015)
D.J. Scheeres, in 36th Annual Lunar and Planetary Science Conference, ed. S. Mackwell, E. Stansbery, p.
1919 (2005)
E.R.D. Scott, G.J. Taylor, H.E. Newsom, et al., in Asteroids II, ed. R. Binzel, T. Gehrels, M. Matthews
(Tucson: University of Arizona Press), p. 701 (1989)
Z. Sekanina, Astrophys. J. 597, 1237 (2003)
C.D. Slaughter, Astron. J. 74, 929 (1969)
A.J. Steffl, N.J. Cunningham, A.B. Shinn, et al., Icarus 223, 48 (2013)
V. Tenishev, M.R. Combi, M. Rubin, Astrophys. J. 732, 104 (2011)
M.H.H. van Dijk, P.B. Bosma, J.W. Hovenier, Astron. Astrophys. 201, 373 (1988)
R. van Lieshout, M. Min, C. Dominik, Astron. Astrophys. 572, A76 (2014)
A. Vourlidas, R.A. Howard, S.P. Plunkett, et al., Space Sci. Rev. online first, DOI 10.1007/s11214-014-0114-
y (2015)
M.K. Wallis, in Comets, ed. L. L. Wilkening (Tucson: University of Arizona Press), p. 357 (1982)
W. Wamsteker, Astron. Astrophys. 97, 329 (1981)
W. Waniak, Icarus 100, 154 (1992)
A. Weigert, Astron. Nachr. 285, 117 (1959)
P.R. Weissman, Astron. Astrophys. 85, 191 (1980)
P.R. Weissman, M.F. A'Hearn, H. Rickman, L.A. McFadden, in Asteroids II, ed. R. Binzel, T. Gehrels, M.
Matthews (Tucson: University of Arizona Press), p. 880 (1989)
F.L. Whipple, W.F. Huebner, Ann. Rev. Astron. Astrophys. 14, 143 (1976)
J.N. Winn, D.C. Fabrycky, Ann. Rev. Astron. Astrophys. 53, 409 (2015)
Q.-Z. Ye, M.-T. Hui, P.G. Brown, et al., Icarus 264, 48 (2016)
M.E. Zolensky, T.J. Zega, H. Yano, et al., Science 314, 1735 (2006)
|
1903.00723 | 1 | 1903 | 2019-03-02T15:25:41 | The future of stellar occultations by distant solar system bodies: perspectives from the Gaia astrometry and the deep sky surveys | [
"astro-ph.EP"
] | Distant objects in the solar system are crucial to better understand the history and evolution of its outskirts. The stellar occultation technique allows the determination of their sizes and shapes with kilometric accuracy, a detailed investigation of their immediate vicinities, as well as the detection of tenuous atmospheres. The prediction of such events is a key point in this study, and yet accurate enough predictions are available to a handful of objects only. In this work, we briefly discuss the dramatic impact that both the astrometry from the Gaia space mission and the deep sky surveys -- the Large Synoptic Survey Telescope in particular -- will have on the prediction of stellar occultations and how they may influence the future of the study of distant small solar system bodies through this technique. | astro-ph.EP | astro-ph |
The future of stellar occultations by distant solar
system bodies: perspectives from the Gaia astrometry
and the deep sky surveys
J. I. B. Camargo1,2 , J. Desmars3 , F. Braga-Ribas4,2, R. Vieira-Martins1,2 , M.
Assafin5,2, B. Sicardy3, D. B´erard3 , G. Benedetti-Rossi1,2
1Observat´orio Nacional / MCTIC, Rua General Jos´e Cristino 77, 20921-400, Rio de
Janeiro, Brazil
2LIneA, Rua General Jos´e Cristino 77, 20921-400, Rio de Janeiro, Brazil
3LESIA / Observatoire de Paris, CNRS UMR 8109, Universit´e Pierre et Marie Curie,
Universit´e Paris-Diderot, 5 place Jules Janssen, F-92195 Meudon C´edex, France
4Federal University of Technology-Paran´a (UTFPR / DAFIS), Rua Sete de Setembro,
3165, CEP 80230-901, Curitiba, PR, Brazil
5Observat´orio do Valongo / UFRJ, Ladeira do Pedro Antonio 43, RJ 20080-090, Rio de
Janeiro, Brazil
Abstract
Distant objects in the solar system are crucial to better understand the his-
tory and evolution of its outskirts. The stellar occultation technique allows
the determination of their sizes and shapes with kilometric accuracy, a detailed
investigation of their immediate vicinities, as well as the detection of tenuous
atmospheres. The prediction of such events is a key point in this study, and
yet accurate enough predictions are available to a handful of objects only. In
this work, we briefly discuss the dramatic impact that both the astrometry from
the Gaia space mission and the deep sky surveys -- the Large Synoptic Survey
Telescope in particular -- will have on the prediction of stellar occultations and
how they may influence the future of the study of distant small solar system
bodies through this technique.
∗Corresponding author
Email address: [email protected] (J. I. B. Camargo1,2)
https://doi.org/10.1016/j.pss.2018.02.014
c(cid:13)2018. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/
Preprint submitted to Journal of LATEX Templates
July 11, 2021
Keywords: Solar system: transneptunians; stellar occultation; Gaia space
mission; deep sky surveys; big data
1. Introduction
Stellar occultation is a powerful technique that allows the determination
of sizes and shapes of distant solar system objects with kilometric accuracy
[1] [2] [3] [4] [5] [6] [7] (leading to albedos, densities), an investigation of their
5
immediate vicinities [6] [7] [8] (telling about the presence of rings, satellites,
jets), and that may reveal tenuous - down to few nanobars - atmospheres [2] [3]
[6] [9].
Accurate predictions1 of occultation events are the very first step for the
full success in the use of the technique. The relevance of this step is such that
10
its improvement inevitably -- and positively -- affects the future of the study of
distant small solar system bodies through stellar occultations.
Thanks to the astrometry from the Gaia space mission and the deep sky
surveys, a huge advance in this study is closer than ever. More specifically, the
first will provide over 1 billion stars with unprecedented (sub milli- to micro-
15
arcsecond) astrometric accuracy (see, for instance, [10] [11]), while the latter,
like the Large Synoptic Survey Telescope (LSST), will provide images from
which short-term accurate ephemerides2 of faint (down to r ∼ 24.5 in the case
of the LSST [12]) solar system bodies can be determined. As a result, this
will lead us to milli-arcsecond (mas) level - or better - predictions to tens of
20
thousands of TNOs as explained in the next two sections.
1Where and when, on the Earth, an occultation event can be observed.
2Ephemerides whose uncertainties, for 1-3 years after the most recent observation used in
the determination of these ephemerides, are smaller than the angular size of the respective
occulting bodies as seen from the Earth.
2
Table 1: Quantities from Fig. 1, lower panels
Technique
Occultation
Direct imaging
Direct imaging
∆αcosδ
0 (±8)
−1 (±4)
6 (±23) −3 (±10)
22 (±42)
14 (±36)
∆δ Dots
Code Gaia
15
red
8
light blue
15
dark blue
Y
Y
N
Technique: way that positions were determined (from occultation or from direct
imaging); ∆αcosδ and ∆δ: average of the differences in the sense position minus
NIMA. Values between parenthesis are the respective standard deviations; Dots:
number of measurements used to derive the values in the two previous columns;
Code: colour code as given in Fig. 1; Gaia: Gaia-based position? Yes/No.
Angular measurements are in units of mas.
2. The power of Gaia
Figure 1 shows the improvement in accuracy, thanks to Gaia, of the ephemeris
of (10199) Chariklo as determined by our orbit fitting tool NIMA [13]. In that
figure, the most relevant difference between its upper and lower panels is that
25
the first is based on pre-Gaia astrometry, whereas the latter is dominated by
Gaia DR1 [14] -based positions [15] (but see also [16] for a foretaste of the
impressive data that will be delivered by the Gaia Data Release 2 from April
2018).
It should be noted that the uncertainty in the most recent version3 of that
30
ephemeris (lower panels) is significantly smaller than the angular size of Chariklo
throughout 2018. And there are more accurate data and orbits to come! Note
that all positions of Chariklo used here are from ground-based observations
and that Gaia DR1 does not provide positions of small solar system bodies.
Thousands of them, however, will be available in Data Release 2 [16],[17].
35
Currently, orbits determined by NIMA discriminates observational data (po-
3On the date this paper was written.
3
Figure 1: Orbit improvement for Chariklo. Versions 8 (2016/JUN, upper panels) and 13
(2017/JUL, lower panels, the most recent one on the date this paper was written) of Chariklo's
orbit as determined by NIMA. In all plots we have: differences in right ascension and decli-
nation (black curves) in the sense NIMA ephemeris minus the Jet Propulsion Laboratory
(JPL) ephemeris (version: JPL20); 1-σ uncertainty (grey area) of NIMA ephemeris; red
points: differences between positions obtained from stellar occultations and those from the
JPL ephemeris; dark/light blue points: differences between positions obtained from direct
imaging of Chariklo and those from the JPL ephemeris; red vertical segments are provided
for an easier visual correspondence between the dots and the dates associated to them; error
bars represent the standard deviation of the observational residuals from the same night and
same observatory. Red dots, in particular, represent one observation each so that no error bar
is attributed to them. In the upper panels, all positions are based on pre-Gaia astrometric
catalogues. In the lower panels, red and light blue points are now Gaia DR1-based. Some of
these light blue points are a re-reduction of the non Gaia-based dark blue points in the upper
panels. Note how well the Gaia-based positions agree with the orbit. Chariklo and its rings
are also roughly represented in the lower right panel.
4
Figure 2: Logarithm of the weights (σ−2) attributed to the points shown in the lower panels
of Fig. 1. The same colour code is used.
sitions) of solar system bodies between those obtained from direct imaging and
those obtained from occultations. It also discriminates between Gaia- and non
Gaia-based positions. In the lower panels of Fig. 1, Gaia-based positions are
given by the red and light blue points whereas non Gaia-based positions are
40
given by the dark blue points. These discriminations are expressed in terms of
weights.
Figure 2 indicates the different weights attributed to the positions mentioned
above. Astrometric data from an occultation is obtained from the relative posi-
tion of the occulting body with respect to that of the occulted star. This relative
45
position is determined with mas level accuracy (see, for instance, [18]). There-
fore, accurate stellar positions, as those given by the Gaia mission, provide mas
level accurate positions of the occulting (solar system) body. In this context, it
is natural that these points (red dots in Fig. 2) have the largest weights.
Table 1 quantifies the effect of the weighing scheme (see [13] for more details),
50
that results in an orbit heavily dominated by the Gaia-based positions with
emphasis to those from occultations.
5
Figure 3: Same as the lower panels of Fig. 1, but excluding the occultation data (red points)
from the fit.
The impact of the use of Gaia DR1-based positions from occultations can
be seen in Fig. 3, from which the occultation data have been withdrawn. As
compared to the lower panels of Fig. 1, we note not only a non negligible (at
55
least in the context of predictions) difference between both ephemerides to those
dates after that of the last observation but also a considerable increase in the
uncertainty (grey zone). It is also important to note that, adding the occultation
data, the ephemeris better fits the light blue (and also Gaia DR1-based) points,
as expected. In this way, in the absence of occultation data, higher weights to
60
the light blue points could be considered.
3. The power of LSST
The Minor Planet Center lists, to date, around 2 600 transneptunian objects
(TNOs) and Centaurs. The LSST, whose full science operations are scheduled to
65
begin in 2023, will record the entire visible sky from Cerro Pach´on about twice
a week and observe millions of solar system objects, ∼40 000 TNOs among them
[12].
The survey expects to deliver astrometry accurate to 10 mas per exposure
depending, among others, on the seeing and signal-to-noise ratio. In some cases,
however, mas level accuracies may be required by stellar occultation predictions
6
70
to observe satellites, grazing events by rings, topographic features, and bodies
that may retain atmosphere. Therefore, a more careful astrometry may also be
needed in specific cases and the previous section showed that this is possible.
It should be noted that many thousands of TNOs are expected to have
more than one hundred observations along the ten years of operations of the
75
survey [12]. These observations are crucial to, in association with the astrometry
from Gaia, orbit fitting tools, and careful weighing schemes (e.g. [13]), obtain
accurate enough short-term ephemerides to all those objects.
4. Comments and conclusions
A stellar occultation event is magnitude independent, in the sense that it
80
only needs to record the flux variation of the star in an interval of time that
contains its occultation. Therefore, the technique is suitable to also investigate
the faintest occulting bodies. With Gaia and deep sky surveys we can expect
the need to select events by focusing on, for instance, small groups of objects
with different physical/dynamical features, instead of trying to observe them
85
all. On the other hand, with the increasing amount of successful observations,
we could also envisage a data-driven approach in the study of small solar system
bodies uniquely from the occultations! The full profit of this exciting and quickly
approaching scenario will, among others, rely on accurate predictions, on the
amateur community, and on the availability of networks of small (20-40 cm)
90
telescopes (e.g.
[19]) with fast readout detectors around the world, as well as
the appropriate support to storage and data processing (big data context).
5. Acknowledgments
J.I.B.C. acknowledges CNPq grant 308150/2016-3. The work leading to
these results has received funding from the National Institute of Science and
95
Technology of the e-Universe project (INCT do e-Universo, CNPq grant 465376/2014-
2). The work leading to these results has received funding from the European
Research Council under the European Community's H2020 2014-2020 ERC
7
grant Agreement no 669416 "Lucky Star". The authors acknowledge the com-
ments from the anonymous referees.
100
6. References
References
[1] J. L. Elliot, M. J. Person, C. A. Zuluaga, et al., Size and albedo of Kuiper
belt object 55636 from a stellar occultation, Nature 465 (2010) 897 -- 900.
doi:10.1038/nature09109.
105
[2] B. Sicardy, J. L. Ortiz, M. Assafin, et al., A Pluto-like radius and a high
albedo for the dwarf planet Eris from an occultation, Nature 478 (2011)
493 -- 496. doi:10.1038/nature10550.
[3] J. L. Ortiz, B. Sicardy, F. Braga-Ribas, et al., Albedo and atmospheric
constraints of dwarf planet Makemake from a stellar occultation, Nature
110
491 (2012) 566 -- 569. doi:10.1038/nature11597.
[4] F. Braga-Ribas, B. Sicardy, J. L. Ortiz, et al., The Size, Shape, Albedo,
Density, and Atmospheric Limit of Transneptunian Object (50000) Quaoar
from Multi-chord Stellar Occultations, ApJ 773 (2013) 26. doi:10.1088/
0004-637X/773/1/26.
115
[5] A. R. Gomes-J´unior, B. L. Giacchini, F. Braga-Ribas, et al., Results of
two multichord stellar occultations by dwarf planet (1) Ceres, MNRAS 451
(2015) 2295 -- 2302. arXiv:1504.04902.
[6] J. L. Ortiz, P. Santos-Sanz, B. Sicardy, et al., The size, shape, density and
ring of the dwarf planet Haumea from a stellar occultation, Nature 550
120
(2017) 219 -- 223. doi:10.1038/nature24051.
[7] F. Braga-Ribas, B. Sicardy, J. L. Ortiz, et al., A ring system detected
around the Centaur (10199) Chariklo, Nature 508 (2014) 72 -- 75. arXiv:
1409.7259, doi:10.1038/nature13155.
8
[8] J. L. Ortiz, R. Duffard, N. Pinilla-Alonso, et al., Possible ring material
125
around centaur (2060) Chiron, A&A 576 (2015) A18. arXiv:1501.05911,
doi:10.1051/0004-6361/201424461.
[9] T. Widemann, B. Sicardy, R. Dusser, et al., Titania's radius and an upper
limit on its atmosphere from the September 8, 2001 stellar occultation,
Icarus 199 (2009) 458 -- 476. doi:10.1016/j.icarus.2008.09.011.
130
[10] F. Mignard, Overall Science Goals of the Gaia Mission, in: C. Turon, K. S.
O'Flaherty, M. A. C. Perryman (Eds.), The Three-Dimensional Universe
with Gaia, Vol. 576 of ESA Special Publication, 2005, p. 5.
[11] Gaia Collaboration, T. Prusti, J. H. J. de Bruijne, A. G. A. Brown,
A. Vallenari, C. Babusiaux, C. A. L. Bailer-Jones, U. Bastian, M. Bier-
135
mann, D. W. Evans, et al., The Gaia mission, A&A 595 (2016) A1.
arXiv:1609.04153, doi:10.1051/0004-6361/201629272.
[12] LSST Science Collaboration, P. A. Abell, J. Allison, et al., LSST Science
Book, Version 2.0, (2009) ArXiv e-printsarXiv:0912.0201.
[13] J. Desmars, J. I. B. Camargo, F. Braga-Ribas, et al., A&A 584 (2015) A96.
140
doi:10.1051/0004-6361/201526498.
[14] Gaia Collaboration, A. G. A. Brown, A. Vallenari, et al., A&A 595 (2016)
A2. doi:10.1051/0004-6361/201629512.
[15] L. Lindegren, U. Lammers, U. Bastian, et al., Gaia Data Release 1. Astrom-
etry: one billion positions, two million proper motions and parallaxes, A&A
145
595 (2016) A4. arXiv:1609.04303, doi:10.1051/0004-6361/201628714.
[16] D. Katz, A. G. A. Brown, Gaia: on the road to DR2, (2017) ArXiv e-
printsarXiv:1710.10816.
[17] P. Tanga, SMALL BODY OBSERVATIONS WITH GAIA, ACM2017 -
Abstract Book, (2017) http://acm2017.uy/abstracts/Plenary6.a.3.pdf.
9
150
[18] B. Sicardy, G. Bolt, J. Broughton, T. Dobosz, D. Gault, S. Kerr,
F. B´enard, E. Frappa, J. Lecacheux, A. Peyrot, J.-P. Teng-Chuen-Yu,
W. Beisker, Y. Boissel, D. Buckley, F. Colas, C. de Witt, A. Doressoundi-
ram, F. Roques, T. Widemann, C. Gruhn, V. Batista, J. Biggs, S. Di-
eters, J. Greenhill, R. Groom, D. Herald, B. Lade, S. Mathers, M. Assafin,
155
J. I. B. Camargo, R. Vieira-Martins, A. H. Andrei, D. N. da Silva Neto,
F. Braga-Ribas, R. Behrend, Constraints on Charon's Orbital Elements
from the Double Stellar Occultation of 2008 June 22, AJ 141 (2011) 67.
doi:10.1088/0004-6256/141/2/67.
[19] M. W. Buie, J. M. Keller, L. H. Wasserman, RECON - A new system for
160
probing the outer solar system with stellar occultations, European Plane-
tary Science Congress 10 (2015) EPSC2015 -- 845.
10
|
1801.03991 | 1 | 1801 | 2018-01-11T21:10:05 | Hall Effect in the coma of 67P/Churyumov-Gerasimenko | [
"astro-ph.EP",
"physics.space-ph"
] | Magnetohydrodynamics simulations have been carried out in studying the solar wind and cometary plasma interactions for decades. Various plasma boundaries have been simulated and compared well with observations for comet 1P/Halley. The Rosetta mission, which studies comet 67P/Churyumov-Gerasimenko, challenges our understanding of the solar wind and comet interactions. The Rosetta Plasma Consortium observed regions of very weak magnetic field outside the predicted diamagnetic cavity. In this paper, we simulate the inner coma with the Hall magnetohydrodynamics equations and show that the Hall effect is important in the inner coma environment. The magnetic field topology becomes complex and magnetic reconnection occurs on the dayside when the Hall effect is taken into account. The magnetic reconnection on the dayside can generate weak magnetic filed regions outside the global diamagnetic cavity, which may explain the Rosetta Plasma Consortium observations. We conclude that the substantial change in the inner coma environment is due to the fact that the ion inertial length (or gyro radius) is not much smaller than the size of the diamagnetic cavity. | astro-ph.EP | astro-ph | Hall Effect in the coma of 67P/Churyumov-Gerasimenko
Z. Huang1, G. Tóth1, T. I. Gombosi1, X. Jia1, M. R. Combi1, K. C. Hansen1, N. Fougere1,
Y. Shou1, V. Tenishev1, K. Altwegg2 and M. Rubin2
Received
;
accepted
8
1
0
2
n
a
J
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
9
9
3
0
.
1
0
8
1
:
v
i
X
r
a
1Climate and Space Sciences and Engineering, University of Michigan, Ann Arbor, MI 48109,
USA
2Physikalisches Institut, University of Bern, Bern, Switzerland
-- 2 --
ABSTRACT
Magnetohydrodynamics simulations have been carried out in studying the so-
lar wind and cometary plasma interactions for decades. Various plasma boundaries
have been simulated and compared well with observations for comet 1P/Halley. The
Rosetta mission, which studies comet 67P/Churyumov-Gerasimenko, challenges our
understanding of the solar wind and comet interactions. The Rosetta Plasma Consor-
tium observed regions of very weak magnetic field outside the predicted diamagnetic
cavity. In this paper, we simulate the inner coma with the Hall magnetohydrodynamics
equations and show that the Hall effect is important in the inner coma environment.
The magnetic field topology becomes complex and magnetic reconnection occurs on
the dayside when the Hall effect is taken into account. The magnetic reconnection on
the dayside can generate weak magnetic filed regions outside the global diamagnetic
cavity, which may explain the Rosetta Plasma Consortium observations. We conclude
that the substantial change in the inner coma environment is due to the fact that the ion
inertial length (or gyro radius) is not much smaller than the size of the diamagnetic
cavity.
Subject headings: MHD, comets: individual: Comet 67P/Churyumov-Gerasimenko,
planet-star interactions
1.
Introduction
Cometary magnetospheres is one of the most important topics in planetary science. Because
the nucleus of a comet is usually very small in size ranging from a few hundred meters to tens of
kilometers (e.g., the radius of the nucleus for comet 1P/Halley is about 10 km) and the gravity
is extremely weak (usually considered negligible when simulating the cometary neutral gas and
-- 3 --
plasma), the cometary coma is much larger in size compared to the nucleus itself. For example,
Giotto observed plasma boundaries of comet 1P/Halley starting at roughly 1 Mkm away from
the nucleus (Reme et al. 1986). The cometary magnetosphere resulting from the solar wind
interaction with the coma has some distinct features from the magnetospheres associated with
planets or planetary moons, such as the formation of a diamagnetic cavity. Gombosi (2015)
provided an excellent review of the cometary magnetosphere. A typical cometary magnetosphere
for an active comet near perihelion includes a bow shock which slows down the supersonic solar
wind to subsonic speed (Galeev et al. 1985; Koenders et al. 2013), a diamagnetic cavity inside
which the magnetic field drops to zero (Neubauer et al. 1986; Cravens 1986; Goetz et al. 2016a,b),
a recombination layer which separates the inner shock (which slows down the supersonic cometary
ion outflow to subsonic) and the contact surface (where the solar wind protons cannot penetrate).
One of the primary goals of the Rosetta Plasma Consortium (RPC) was to observe the evolution
of the solar wind and comet interactions. However, due to the close proximity of the Rosetta
spacecraft to the nucleus, RPC was not able to observe the bow shock. Also the recombination
layer and contact surface have not been clearly identified to date. Based on RPC observations,
Mandt et al. (2016) reported plasma boundaries separating an inner region and an outer region
and they concluded the observed plasma boundaries are an ion-neutral collisionopause boundary,
which has not been predicted by previous numerical simulations (Rubin et al. 2015; Koenders et al.
2015; Huang et al. 2016). In addition to this unpredicted boundary, the magnetic field observed
by RPC is also unexpected: the diamagnetic 'cavity' (Goetz et al. 2016a,b) was observed much
farther away than the predicted locations (Rubin et al. 2015; Koenders et al. 2015; Huang et al.
2016).
Lots of effort has been made in numerical simulations to understand the solar wind and comet
interactions. There are two major approaches in simulating the cometary environment: the fluid
approach (Gombosi et al. 1996; Hansen et al. 2007; Rubin et al. 2014, 2015; Huang et al. 2016) and
the hybrid approach (Bagdonat & Motschmann 2002; Koenders et al. 2015; Wedlund et al. 2017).
-- 4 --
In a fluid approach, the plasma is treated as fluids and governed by the magnetohydrodynamic
(MHD) equations. The fluid approach is an accurate description of the macroscopic quantities
of the plasma when the Knudsen number (Kn = λ
l , where λ is the mean free path and l is the
characteristic length scale of the flow) is much smaller than unity. On the other hand, the hybrid
approach simulates ions as individual particles and electrons as a fluid. The hybrid approach
can capture the kinetic features of the plasma and works well also for large Knudsen numbers.
Compared to the fluid approach, the hybrid approach is very computationally expensive and is
usually limited to a small simulation domain. Recently, a third approach to simulate the solar
wind and comet interactions was developed by Deca et al. (2017) with a fully kinetic code, which
treats both ions and electrons as particles. They simulated comet CG at 3 au and showed that
their simulations agreed well with RPC observations at that heliocentric distance. However, their
code does not include any chemical reactions and collisions between particles, which makes it not
applicable to comets near perihelion, where chemical reactions are important. Also their model is
limited to a small domain that does not include the bow shock.
There have been extensive discussions about whether fluid codes can properly simulate
cometary environments or those of other planets/moons without an intrinsic magnetic field, in
which case the ion gyro radius is larger than the length scale of interaction regions. We argue
that fluid codes are not limited exclusively by the requirement that the length scale must be much
larger than the ion gyro radius, because collisions among different particles as well as chemical
reactions may reduce the ion kinetic effects arising from the gyro motion. It is true that single
fluid ideal MHD models cannot capture any ion gyration effects compared to a hybrid simulation
(Hansen et al. 2007). But when fluid simulations take into account different fluids with different
velocities as well as collisions among them, multi-fluid MHD models are capable of capturing
some important ion kinetic behaviors. For example, Rubin et al. (2014) showed that their
multi-fluid MHD model is capable of resolving the gyration of different ion fluids with reasonably
good agreement with what has been predicted by a hybrid model (Müller et al. 2011). The major
-- 5 --
discrepancy lies in the very inner coma region with striations/filaments in the cometary ion density
(Koenders et al. 2015). On the other hand, multi-fluid simulations for other planets/moons without
an intrinsic magnetic field have demonstrated that the simulation results agree well with in-situ
plasma observations (Najib et al. 2011; Bougher et al. 2015; Ma et al. 2011). Hybrid models
certainly provide a better description of the plasma environment near the comet nucleus. However,
it is much more expensive or nearly impossible computationally to run hybrid models with a
grid resolution comparable to those of fluid models on a large enough domain to properly set up
the outer boundary conditions and to resolve the details of a diamagnetic cavity formed close to
the nucleus. With a relatively coarse grid typically used in hybrid models, effects of numerical
diffusion are expected to be much stronger than in fluid models and in such a case, the evolution of
the magnetic field may not be properly described. As described below, the multi-fluid Hall MHD
model presented in this paper represents another step in further resolving kinetic effects with fluid
simulations.
One of the imperfections of previous MHD models applied in cometary studies is that
the Hall effect is usually not taken into account. The Hall effect describes the relative speed
(current) between ions and electrons and appears in the generalized Ohm's law. This current may
affect the magnetic field evolution in the system if the Hall effect is taken into account in the
induction equation. The Hall effect is important in magnetic reconnection studies as Hall MHD
is the minimal modification of resistive MHD that can reproduce the fast reconnection process
(Birn et al. 2001), partially due to the strong current near the reconnection null point. In the
cometary magnetosphere, the diamagnetic cavity is a unique feature that other planets/moons do
not have. As the magnetic field drops to zero in a short distance, there must be strong currents
along the diamagnetic cavity boundary. How these currents affect the inner coma environment is
still unknown. In this paper, we simulate the inner coma environment with Hall MHD equations
and show that the Hall effect is important in the inner coma and the classical plasma boundaries
obtained by previous models need to be revisited. The detailed model description can be found in
the appendix.
-- 6 --
2. The Hall MHD model
Our Hall MHD model is an extension of the multifluid model developed by Huang et al.
(2016). In the following equations, mass density, velocity vector, pressure, the identity matrix and
the adiabatic index are denoted by symbols ρ, u, p, I and γ, respectively. The cometary neutral
gas, the ions (cometary and solar wind) and the electrons are denoted by subscripts n, s and e,
respectively. The symbol Z denotes the ion charge state while the symbol e is for the unit charge.
There are four fluids in the model. One fluid describes the cometary neutral gas with the
Euler equations:
+ ∇ · (ρnun) =
∂ ρn
δρn
∂t
δt
+ ∇ · (ρnunun + pnI) =
∂ ρnun
∂t
δt
+ ∇ · (pnun) + (γn − 1)pn(∇ · un) =
δρnun
∂pn
∂t
δpn
δt
and the other two fluids describe the cometary ions and the solar wind protons with the multifluid
MHD equations, which are solved individually for both fluids:
(1a)
(1b)
(1c)
(2a)
(2b)
(2c)
∂ ρsus
∂t
δρs
δt
+ ∇ · (ρsus) =
∂ ρs
∂t
+ ∇ · (ρsusus + psI)
− Zse ρs
ms
(E + us × B) =
δρsus
δt
+ ∇ · (psus) + (γs − 1)ps(∇ · us) =
δps
δt
∂ps
∂t
charge neutrality in the plasma, the electron number density can be obtained as ne =
For the electrons, we do not specify the continuity and momentum equations. Assuming
s=ions Zsns.
-- 7 --
s=ions Zsnsus
ne
The electron velocity ue is obtained from ue = u+ + uH, where u+ is the charge averaged ion
velocity (u+ =
j = (1/µ0)∇ × B). The electron pressure in the system is described by Equation 3:
) and uH is the Hall velocity (uH = − j
nee, where j is the current density
∂pe
∂t
+ ∇ · (peue) + (γe − 1)pe(∇ · ue) =
δpe
δt
(3)
We use Equations (1) - (3) to describe the behavior and interactions of different fluids (the
cometary neutral gas, the cometary ions, the solar wind protons, and the electrons) in the system.
Ionization (photo-ionization and electron impact ionization) of the cometary neutral gas, charge
exchange between neutrals and ions, collisions (elastic and inelastic) between different fluids, and
recombination are all taken into account in simulating the cometary environment and they appear
as source terms in the right hand side of Equations (1) - (3). We apply the same source terms as
Huang et al. (2016). The stiffness of the source terms may limit the time step, so a point-implicit
algorithm (Tóth et al. 2012) is applied to evaluate these terms.
The electric and magnetic fields are also needed to solve the multifluid equations. The electric
field is derived from the electron momentum equation if the inertial terms are assumed to be zero
(due to the small electron mass):
E = −ue × B − 1
nee
∇pe
The magnetic field is obtained from the induction equation:
∂B
∂t
= −∇ × E
(4)
(5)
We solve Equations (1) to (5) on a 3D block adaptive grid with the BATS-R-US (Block-
Adaptive Tree Solarwind Roe-type Upwind Scheme) code (Powell et al. 1999; Tóth et al. 2012).
The beauty of the adaptive grid is that we can resolve different length scales in the system, so that
the simulation can resolve the nucleus while modeling the global scales. In the comet CG case, the
radius of the nucleus is about 2 km, the global diamagnetic cavity is reported to be about 100 km,
-- 8 --
and the bow shock is expected to be at about 8,000 km upstream of the nucleus (Rubin et al. 2015;
Koenders et al. 2015; Huang et al. 2016). In our simulation, the smallest cell is located near the
nucleus with the size of about 0.12 km and the largest cell is located near the outer boundary
with the size of about 31,250 km, which requires 18 levels of refinements (each refinement level
increases the resolution by a factor of two) in the domain. We use the Cometocentric Solar
Equatorial (CSEQ) frame in the simulation. In this frame, +x points toward the Sun, the z axis
contains the solar rotation axis, and the y axis is orthogonal to the x and z axes. The solar wind is
considered to move along the -x direction with the interplanetary magnetic field points in the +y
direction at the upstream boundary. The simulation box is within ±106 km in the x direction and
±0.5 × 106 km in both y and z directions. We specify boundary conditions the same way as Huang
et al. (2016) at the edge of the simulation box (outer boundary) as well as at the nucleus surface
(inner boundary).
In the present study, the cometary neutral gas is limited to water molecules with the specific
3 and the corresponding cometary ions are H2O+ with the same γ. γ = 5
heat ratio (γ) of 4
3 is
applied for the solar wind protons as well as electrons. An idealized spherical comet with the
neutral gas outflow driven by the solar illumination (hereafter illuminated sphere) seems to be the
minimum requirement not to lose important asymmetrical features in the inner coma (Huang et al.
2016), so we apply this nucleus condition at the inner boundary, which is the same as Case 2 in
Huang et al. (2016). We apply the same input parameters listed in Tables 2 and 3 in Huang et al.
(2016). We first run the multifluid model in steady state mode without the Hall effect to reach a
steady state. We then introduce the Hall effect at t = 0 and run the model in time-dependent mode
to investigate the evolution of the inner coma.
-- 9 --
Fig. 1. -- The Hall MHD simulation results. The left columns plot the y=0 plane while the right
columns plot the z=0 plane. The upper panels are for the cometary ion density while the lower
panels show the magnetic field magnitude.
3. Simulation results
Figure 1 shows the multifluid Hall MHD simulation results in the inner coma region (within
400 km of the nucleus) with an illuminated sphere at t = 600 s. While the simulation results
preserve symmetries about the y axis in the z=0 plane, they show pronounced asymmetries in the
2.02.53.03.54.04.55.0 log H2O+ density [cm-3] -400-2000200400Z [km] 01020304050 Magnetic field magnitude [nT]-400-2000200400X [km]-400-2000200400Z [km] 2.02.53.03.54.04.55.0 log H2O+ density [cm-3] -400-2000200400Y [km] 01020304050 Magnetic field magnitude [nT]-400-2000200400X [km]-400-2000200400Y [km] -- 10 --
y = 0 plane. As a comparison, we reproduce the multifluid simulation results without the Hall
effect in the same region with an illuminated sphere for the same input parameters (Case 2 in
Huang et al. (2016)) in Figure 2. In Huang et al. (2016), the size of the diamagnetic cavity and the
location of the contact surface agreed well with previous MHD simulations (Rubin et al. 2015) and
hybrid simulations (Koenders et al. 2015). However, when the Hall effect is introduced, the ion
pile-up region (with light yellow color) in the upper panel is distorted in the y=0 plane and looks
completely different from the upper panel in Figure 2, where the distribution is symmetric about
the z axis. Some surface wave structures, which might be associated with the Kelvin-Helmoltz
(hereafter K-H) instabilities reported in Rubin et al. (2012), can also be found in the upper panel in
Figure 1.
The magnetic field topology in the bottom panels in Figure 1 is completely different from
the bottom panels in Figure 2. When the model does not include the Hall effect, the diamagnetic
cavity (bottom panels in Figure 2) is an isolated region and the magnetic field pile-up region is
just upstream of the diamagnetic cavity. When the Hall effect is introduced, the magnetic field
configuration becomes more complex and besides the 'global' diamagnetic cavity, regions of very
weak magnetic field (less than 10 nT) can also be found in the lower right corner in the y=0 plane
in the bottom left panel in Figure 1. We suggest that it is the J × B force discussed in the next
paragraph that changes the magnetic field configuration in the inner coma region. The magnetic
field pile-up region in the y=0 plane is shifted and is not located in the same region as in Figure 2.
In the z=0 plane (the bottom right panel in both Figure 1 and Figure 2), the diamagnetic cavity
looks more or less the same between the two simulations. The biggest difference lays in the
magnetic pile-up region. In Figure 1, only two small magnetic field pile-up regions are found
outside the diamagnetic cavity while in Figure 2, the magnetic pile-up region is a single region.
It is quite surprising that the simulated magnetic field changes so dramatically when the Hall
effect is taken into account in the induction equation. Besides, the Hall MHD simulation does
-- 11 --
Fig. 2. -- A reproduction of the simulation results from Huang et al. (2016). Multi-fluid MHD
results without the Hall effect to be compared with Figure 1.
2.02.53.03.54.04.55.0 log H2O+ density [cm-3] -400-2000200400Z [km] 0102030405060 Magnetic field magnitude [nT]-400-2000200400X [km]-400-2000200400Z [km] 2.02.53.03.54.04.55.0 log H2O+ density [cm-3] -400-2000200400Y [km] 0102030405060 Magnetic field magnitude [nT]-400-2000200400X [km]-400-2000200400Y [km] -- 12 --
Fig. 3. -- The upper panel shows the cometary ion density with their streamlines on the y=0 plane.
The bottom panels plots the z component of the J×B force density (in the unite of 1× 10−12 N/m3)
on the y=0 plane.
2.02.53.03.54.04.55.0 log H2O+ density [cm-3] -400-2000200400Z [km] -400-2000200400Z [km] -400-2000200400Z [km] -400-2000200400Z [km] -20-1001020 (JxB)z [pN/m3]-400-2000200400X [km]-400-2000200400Z [km]-400-2000200400X [km]-400-2000200400Z [km]-400-2000200400X [km]-400-2000200400Z [km]-400-2000200400X [km]-400-2000200400Z [km] -- 13 --
not have a steady state solution despite the fixed upstream solar wind conditions. The online
movie ('inner_coma_movie.mp4') shows the evolution of the cometary ion density (with velocity
streamlines) and the magnetic field between t = 421 s and t = 600 s. In the movie, the cometary
ions move in the negative z direction. To illustrate this, we plot a snapshot of the cometary ion
density with velocity streamlines at t = 600 s in the upper panel of Figure 3. This motion of the
cometary ions can be explained by the J × B force, which is plotted in bottom panel of Figure 3.
This figure shows that along the global diamagnetic cavity boundary, the J× B force has a negative
z component, which acts to move the cometary ions in the negative z direction.
Another important new observation from the Hall MHD simulation is the formation of
the weak magnetic field regions in the lower right corner in Figure 1. These structures appear
as quasi-periodic structures in the online movie ('inner_coma_movie.mp4'). We provide a
best estimate of the periods ranging from 10 s to 50 s, which are a combination of different
harmonic periods, based on the evolution of the magnetic pile-up regions in the online movie
('inner_coma_movie.mp4'). The periods depend on many factors, e.g., the plasma flow speed
compared to the Alfvén speed, the strength of the currents as well as the direction and magnitude
of the J × B force. It is impossible to calculate the exact periods in such a complex case. To
investigate how these weak magnetic field regions form, we have examined the evolution of the
magnetic field topology in 3D, which is animated in another two online movies (with different
view angles, 'reconnection_movie_view1.avi' and 'reconnection_moive_view2.avi'). In the
plasma, the magnetic field is approximately frozen into the electron fluid. Both the cometary ions
in the coma and the electrons move in the negative z direction (with the velocities separated by
the currents). As the magnetic field moves with the electrons, the magnetic field is then draped
in the negative z direction. A recent hybrid simulation by Koenders et al. (2016) also showed
the draping signatures for comet CG at 2.0 au. In the Hall MHD simulation, the draping of the
magnetic field lines forms a configuration that favors magnetic reconnections. The online movies
only animate 15 s of the evolution (between t = 425 s and 440 s), but they clearly show how the
-- 14 --
(cid:113) 1
magnetic field reconnect and forms magnetic flux ropes. The magnetic reconnection reduces
the magnetic field magnitude and create the weak magnetic field regions. Figure 4 plots the 3D
magnetic field configuration. Magnetic reconnections are expected to occur where the magnetic
field lines bend strongly, denoted by an 'X' mark on the figure. As magnetic reconnections occur,
outflow is expected at the magnetic null point with opposite directions. Figure 5 confirms that the
plasma moves oppositely on the two sides. The outflow speed is close to the Alfvén speed near the
reconnection regions, which is in the order of 1 km/s.
Why does the Hall effect matter in the inner coma of comet CG? We argue that it is because
the scale of the diamagnetic cavity is comparable to the ion inertial length (di = mi
ρi µ0 , where
qi
mi is the ion mass, qi is the ion charge, ρi is the ion mass density, µ0 is the magnetic permeability
of vacuum) and the ion gyro radius (ri = vth,imi
qi B , where vth,i is the ion thermal speed, B is the
magnetic field magnitutde). In the inner coma, the cometary ions dominate, so the ion inertial
length and the gyro radius for the cometary ions are responsible for the physical processes. Figure
6 plots the ion inertial length and the gyro radius for the cometary ions. The ion inertial length is
slightly less than 10 km in the ion pile up region, and it is in the order of 100 km outside the global
diamagnetic cavity and the ion pile-up region. The gyro radius is very large where the magnetic
field is small. Except in the weak magnetic field regions, the gyro radius has similar distributions
as the ion inertial length. As the size of the global diamagnetic cavity is about 100 km, the ion
inertial length and the ion gyro radius are not much smaller than the global diamagnetic cavity.
Dorelli et al. (2015) showed that Hall currents within the magnetopause and magnetotail current
sheets have a significant impact on the global structure of Ganymede's magnetosphere, because
the magnetopause standoff distance is not much larger (order of 10) than the ion inertial length.
In our case, the ratio of the size of the diamagnetic cavity and the ion inertial length (or gyro
radius) is in order of 10 or less. We put forward an argument that if the ion inertial length (or gyro
radius) is not much smaller than the characteristic length of the magnetosphere (the diamagnetic
cavity in our case, the magnetopause standoff distance in Ganymede's magnetosphere), Hall MHD
-- 15 --
Fig. 4. -- A 3D views of the magnetic field lines. The contour shows the magnetic field magnitude
at the y=0 plane. Magnetic reconnection is expected to occur near the red cross mark.
-- 16 --
simulations are necessary to capture the correct global structure of the magnetosphere.
4. Summary and Discussions
In this work, we performed a multi-fluid Hall MHD simulation to study the cometary plasma
environment in the inner coma region of comet 67P/Churyumov-Gerasimenko. With the same
model set up as Huang et al. (2016), the Hall MHD simulation shows a very different picture: the
inner coma is no longer symmetric and low magnetic field regions can form outside the global
diamagnetic cavity and the solution is time-dependent.
The only difference between the Hall MHD simulation and the classical MHD simulations by
Huang et al. (2016) is that the Hall velocity term is considered in the magnetic induction equation,
which means that the current can affect the evolution of the magnetic field. It is well known that
the Hall effect is important in magnetic reconnections (Birn et al. 2001), partially due to relative
weak magnetic field and strong currents near the magnetic null point. Hall MHD simulations of
the magnetospheres of planets and moons typically do not show significant differences compared
to the classical MHD simulations except in the regions where magnetic reconnections occur like
the dayside magnetopause or the nightside magnetotail. Dorelli et al. (2015) reported that Hall
effect is important in Ganymede's magnetosphere because the magnetopause standoff distance is
in the order of 10 times larger than the ion inertial length. One would not expect the Hall MHD
simulations dramatically to change the simulated inner coma environment for a comet because
magnetic field reconnections have only been reported on the nightside (Huang et al. 2016), but our
simulations show that in fact the results change dramatically.
The diamagnetic cavity is a unique feature in the cometary environment, which is not shared
by other planets or moons in the solar system, and it has received lots of attention since the Giotto
mission (Neubauer et al. 1986; Cravens 1986; Goetz et al. 2016a,b; Huang et al. 2016; Madanian
-- 17 --
Fig. 5. -- The Uy component for the cometary ion velocity at the surface close to the mag-
netic reconnection surface, which is defined by 3 points: [480, 0, 0] km, [500, 20, 33.5] km and
[460, 0, −33.5] km. The cometary ions move to the +y direction on the right side while they move
to the -y direction on the left side, near the magnetic null point denoted by the red cross mark, as
indicated by the two black arrows.
-- 18 --
Fig. 6. -- The ion inertial length and the ion gyro radius for the cometary ion within 400 km of the
nucleus on the y=0 plane.
0.00.51.01.52.02.53.0 log (di) [km] -400-2000200400Z [km] 0.00.51.01.52.02.53.0 log (ri) [km]-400-2000200400X [km]-400-2000200400Z [km] -- 19 --
et al. 2017). However, it has not been realized that currents along the diamagnetic cavity boundary
may change the global structure of the inner coma. In the comet CG case, as the ion inertial length
(or gyro radius) is not much smaller than the size of the global diamagnetic cavity, the Hall effect
plays an important role in the evolution of the cometary plasma environment in the inner coma
region, which is confirmed by our Hall MHD simulation. The situation might be different for a
much more active comet. For example, the size of the diamagnetic cavity for comet 1P/Halley is
about 4500 km (Neubauer et al. 1986; Cravens 1986), which will need to be compared with the ion
inertial length (or gyro radius) to see whether the Hall effect is important there.
The most important feature from the Hall MHD simulations is that there can be dayside
magnetic reconnection, which can create weak magnetic field regions outside the global
diamagnetic cavity. One of the most puzzling observations from the Rosetta Plasma Consortium
(RPC) is that the magnetometer observed weak magnetic field at a distance much farther away
than the predicted diamagnetic cavity. Goetz et al. (2016a,b) explained the weak magnetic field
observations as K-H instabilities propagating along the cavity boundary and Huang et al. (2016)
explained them as short-lived enhanced electron pressure along magnetic field lines. The Hall
MHD simulation may provide a third option, magnetic field reconnection on the dayside. Further
investigation and data comparison is necessary, but at this point, we refer this to future studies.
Acknowledgements
This work was supported by contracts Jet Propulsion Laboratory no. 1266313 and no.
1266314 from the US Rosetta Project and NASA grant NNX14AG84G from the Planetary
Atmospheres Program.
The authors would like to thank the ROSINA team for supporting this research. The authors
also thank the ESA Rosetta team for providing the opportunities to study this unique comet and
-- 20 --
their continuous support.
The authors would like to acknowledge the following high-performance computing resources:
Yellowstone (ark:/85065/d7wd3xhc), provided by NCAR's Computational and Information
Systems Laboratory, sponsored by the National Science Foundation; Pleiades, provided by the
NASA Supercomputer Division at Ames; and Extreme Science and Engineering Discovery
Environment (XSEDE), supported by National Science Foundation grant number ACI-1053575
-- 21 --
REFERENCES
Bagdonat, T., & Motschmann, U. 2002, Earth Moon and Planets, 90, 305
Birn, J., Drake, J. F., Shay, M. A., et al. 2001, J. Geophys. Res., 106, 3715
Bougher, S., Jakosky, B., Halekas, J., et al. 2015, Science, 350, 0459
Cravens, T. E. 1986, in ESA Special Publication, Vol. 250, ESLAB Symposium on the Exploration
of Halley's Comet, ed. B. Battrick, E. J. Rolfe, & R. Reinhard, 241 -- 246
Deca, J., Divin, A., Henri, P., et al. 2017, Phys. Rev. Lett., 118, 205101
Dorelli, J. C., Glocer, A., Collinson, G., & Tóth, G. 2015, Journal of Geophysical Research (Space
Physics), 120, 5377
Galeev, A. A., Cravens, T. E., & Gombosi, T. I. 1985, ApJ, 289, 807
Goetz, C., Koenders, C., Richter, I., et al. 2016a, A&A, 588, A24
Goetz, C., Koenders, C., Hansen, K. C., et al. 2016b, MNRAS, 462, S459
Gombosi, T. I. 2015, Physics of Cometary Magnetospheres (John Wiley & Sons, Inc), 169 -- 188
Gombosi, T. I., De Zeeuw, D. L., Häberli, R. M., & Powell, K. G. 1996, J. Geophys. Res., 101,
15233
Hansen, K. C., Bagdonat, T., Motschmann, U., et al. 2007, Space Sci. Rev., 128, 133
Huang, Z., Tóth, G., Gombosi, T. I., et al. 2016, MNRAS, 462, S468
Huang, Z., Tóth, G., Gombosi, T. I., et al. 2016, Journal of Geophysical Research: Space Physics,
n/a, 2015JA022333
-- 22 --
Koenders, C., Glassmeier, K.-H., Richter, I., Motschmann, U., & Rubin, M. 2013, Planet. Space
Sci., 87, 85
Koenders, C., Glassmeier, K.-H., Richter, I., Ranocha, H., & Motschmann, U. 2015, Planet. Space
Sci., 105, 101
Koenders, C., Goetz, C., Richter, I., Motschmann, U., & Glassmeier, K.-H. 2016, MNRAS, 462,
S235
Ma, Y. J., Russell, C. T., Nagy, A. F., et al. 2011, Journal of Geophysical Research (Space Physics),
116, A10213
Madanian, H., Cravens, T. E., Burch, J., et al. 2017, AJ, 153, 30
Mandt, K. E., Eriksson, A., Edberg, N. J. T., et al. 2016, MNRAS, 462, S9
Müller, J., Simon, S., Motschmann, U., et al. 2011, Computer Physics Communications, 182, 946
Najib, D., Nagy, A. F., Tóth, G., & Ma, Y. 2011, Journal of Geophysical Research (Space Physics),
116, 5204
Neubauer, F. M., Glassmeier, K. H., Pohl, M., et al. 1986, Nature, 321, 352
Powell, K. G., Roe, P. L., Linde, T. J., Gombosi, T. I., & de Zeeuw, D. L. 1999, Journal of
Computational Physics, 154, 284
Reme, H., Sauvaud, J. A., D'Uston, C., et al. 1986, Nature, 321, 349
Rubin, M., Hansen, K. C., Combi, M. R., et al. 2012, Journal of Geophysical Research (Space
Physics), 117, A06227
Rubin, M., Koenders, C., Altwegg, K., et al. 2014, icarus, 242, 38
-- 23 --
Rubin, M., Jia, X., Altwegg, K., et al. 2015, Journal of Geophysical Research (Space Physics),
120, 3503
Tóth, G., van der Holst, B., Sokolov, I. V., et al. 2012, Journal of Computational Physics, 231, 870
Wedlund, C. S., Alho, M., G., G., & et al. 2017, Astronomy & Astrophysics
This manuscript was prepared with the AAS LATEX macros v5.0.
|
1508.06913 | 1 | 1508 | 2015-08-27T16:07:09 | Failure modes and conditions of a cohesive, spherical body due to YORP spin-up | [
"astro-ph.EP"
] | This paper presents transition of the failure mode of a cohesive, spherical body due to YORP spin-up. On the assumption that the distribution of materials in the body is homogeneous, failed regions first appearing in the body at different spin rates are predicted by comparing the yield condition of an elastic stress in the body. It is found that as the spin rate increases, the locations of the failed regions move from the equatorial surface to the central region. To avoid such failure modes, the body should have higher cohesive strength. The results by this model are consistent with those by a plastic finite element model. Then, this model and a two-layered-cohesive model first proposed by Hirabayashi et al. are used to classify possible evolution and disruption of a spherical body. There are three possible pathways to disruption. First, because of a strong structure, failure of the central region is dominant and eventually leads to a breakup into multiple components. Second, a weak surface and a weak interior make the body oblate. Third, a strong internal core prevents the body from failing and only allows surface shedding. This implies that observed failure modes may highly depend on the internal structure of an asteroid, which could provide crucial information for giving constraints on the physical properties. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- 10 (2002)
Preprint 11 June 2021
Compiled using MNRAS LATEX style file v3.0
Failure modes and conditions of a cohesive, spherical body
due to YORP spin-up
Masatoshi Hirabayashi1(cid:63)
1Aerospace Engineering Sciences, 429 UCB, University of Colorado, Boulder, CO 80309-0429 United States
Accepted 2015 August 26. Received 2015 August 26; in original form 2015 August 15
ABSTRACT
This paper presents transition of the failure mode of a cohesive, spherical body due
to YORP spin-up. On the assumption that the distribution of materials in the body
is homogeneous, failed regions first appearing in the body at different spin rates are
predicted by comparing the yield condition of an elastic stress in the body. It is found
that as the spin rate increases, the locations of the failed regions move from the
equatorial surface to the central region. To avoid such failure modes, the body should
have higher cohesive strength. The results by this model are consistent with those by
a plastic finite element model. Then, this model and a two-layered-cohesive model first
proposed by Hirabayashi et al. are used to classify possible evolution and disruption
of a spherical body. There are three possible pathways to disruption. First, because of
a strong structure, failure of the central region is dominant and eventually leads to a
breakup into multiple components. Second, a weak surface and a weak interior make
the body oblate. Third, a strong internal core prevents the body from failing and only
allows surface shedding. This implies that observed failure modes may highly depend
on the internal structure of an asteroid, which could provide crucial information for
giving constraints on the physical properties.
Key words: minor planets, asteroids: general -- protoplanetary discs -- methods:
analytical -- methods: numerical
1 INTRODUCTION
Recent observations have shown that rotational disruption
may be common in our system. P/2013 R3 has broken into
multiple components (Jewitt et al. 2014), 62412 has had
dust tails from its nucleus (Sheppard & Trujillo 2015), and
P/2012 F5 (GIBBS) has ejected dust and fragments (Drahus
et al. 2015). They are considered to be candidates that have
failed due to fast rotation, implying that rotational disrup-
tion may be diverse.
The YORP effect, which highly depends on the obliq-
uity and shape (Scheeres 2007; Statler 2009; Cotto-Figueroa
et al. 2015), may be one of the primary factors that accel-
erate/decelerate the spin of an asteroid with a few hundred
meters in diameter (Rubincam 2000). We have to mention
that small impacts are also capable of changing the spin
rate of an asteroid quasi-statically (Marzari et al. 2011).
Spin up/down by this effect may be more critical to aster-
oidal evolution in the solar system (Pravec et al. 2006) and
in the debris disk around a white dwarf (Veras et al. 2014)
than ever thought. A remarkable aspect of fast rotating as-
teroids is that observed asteroids spinning at spin periods
(cid:63) E-mail: [email protected]
c(cid:13) 2002 The Authors
shorter than 3.5 hours are all spheroidal (personal communi-
cation with Patrick Taylor, 2015). For example, 1999 KW4
Alpha, which is rotating with a spin period of 2.76 hours,
is a quasi-spherical body with ridges at its equator (Ostro
et al. 2006). The similar features were also seen for the cases
of 1950 DA (Busch et al. 2007) and 1994 CC Alpha (Bro-
zovi´c et al. 2011). This may tell us that fast rotating, quasi-
spherical asteroids are fairly common in the solar system. It
is also important to mention that rotational disruption by
the YORP effect would lead to complex dynamical evolution
(Jacobson & Scheeres 2011).
Such analytical and observational evidence of quasi-
static spin-up by the YORP effect has been giving rise to the
question about evolution and disruption of quasi-spherical
objects. Analytical works have contributed to better under-
standing of failure mechanisms of a uniformly rotating ellip-
soid (Dobrovolskis 1982; Davidsson 2001; Holsapple 2001,
2004; Sharma 2013). Numerical investigations have been ca-
pable of dynamically modeling shape reconfigurations by the
YORP spin-up (Walsh et al. 2008, 2012; S´anchez & Scheeres
2012). These studies gave scientific insights into reshaping
processes of ellipsoidal and quasi-spherical objects affected
by material friction. Significant progress was made by Walsh
et al. (2008) who first succeeded in explaining the formation
2
Hirabayashi
of equatorial ridges on a quasi-spherical body by using a
hard-sphere discrete element code. Also, S´anchez & Scheeres
(2012) used a soft-sphere discrete element code to analyze
the failure modes of cohesionless ellipsoids having different
friction. An ideal shape of an initially spherical body affected
by self-gravity and rotation was numerically and analyti-
cally investigated by considering the angle of repose (Harris
et al. 2009; Scheeres 2015). Regardless of these efforts, the
reshaping process of a quasi-spherical body is still poorly un-
derstood. Specifically, how does the internal structure con-
tribute to evolution and disruption due to quasi-static spin-
up? How does the stress field have an effect on failure at
severe rotation conditions? How does cohesive strength help
an asteroid remain without structural damage?
The motivation of this study comes from earlier studies
about the failure conditions of a quasi-spherical body spin-
ning fast. Assuming a homogeneous structure, Hirabayashi
& Scheeres (2015) found that 1950 DA may mainly experi-
ence failure of its central region, but should not have surface
failure at the current spin period, 2.1 hours. This result in-
dicates that landslides may not make equatorial ridges on
a quasi-spherical asteroid, but inelastic deformation due to
strong compression along the spin axis is a likely driver for
it. Hirabayashi et al. (2015) confirmed that at high spin
rates the central region certainly reaches the yield first and
alternatively proposed that to have surface failure and mass
shedding driven by it a spherical body must have a strong
core. These studies imply that the internal structure plays
a prime role in evolution and disruption of a quasi-spherical
object. The issue in their works, however, is that they only
considered faster spin cases.
The main objective in this paper is to infer the de-
tailed failure mode of a quasi-spherical object at different
spin states and to visualize its evolution and disruption. To
do so we consider an asteroid to be perfectly sphere and an-
alyze its failure modes and conditions by applying a contin-
uum mechanics approach. Theoretical and numerical studies
are employed here. For the theoretical investigation, ana-
lytical solutions of the stress field by linear elasticity are
used to predict when and where failure occurs in a spher-
ical body. For the numerical study, a plastic finite element
model (FEM) developed by Hirabayashi & Scheeres (2015)
is employed to see the detailed conditions for failure of the
body.
It should be noted that the analytical study here di-
rectly applies the technique by Dobrovolskis (1982). He in-
troduced a quantity of shear and pressure called "duress"
to investigate the internal structures of the Martian moons,
Phobos and Deimos, and several asteroids. According to his
study, Phobos should have the maximum shear stress at the
central region at the current orbital and rotational config-
urations, and could only have thrust faults on its surface.
He also studied the fate of Phobos, and found that the tidal
force will have a primary effect on its failure once it reaches
a distance of 6700 km from the center of Mars. We extend
his pioneering work to investigate the case in which a spheri-
cal object changes its stress field due to quasi-static spin-up.
Specifically, we focus on how possible failure modes can shift
in the body by considering the Drucker-Prager yield crite-
rion, a pressure-shear dependent yield condition.
This paper is organized as follows. First, using linear
elasticity, we establish a stress field in a spherical body as
a function of a spin rate. Second, the yield conditions of
the surface region and the central region are analytically
investigated to predict the locations of the failed regions that
appear first. Third, we use a plastic FEM to evaluate the
analytical results. Finally, possible evolution and disruption
scenarios of a spherical body are proposed by taking account
of a two-layered-cohesive model by Hirabayashi et al. (2015).
2 MODELING OF A SPHERICAL ASTEROID
We begin our analysis with modeling of a spherical asteroid
that experiences the YORP spin-up. A spherical object is
supposed to have a homogeneous structure. Although the
model used is the same as Hirabayashi et al. (2015), we do
not explicitly consider heterogeneity in the body as our focus
is on determining the critical values of material properties
with which local regions fail structurally. The density, the
radius and the gravitational constant are defined as ρ, R and
G, respectively. The following discussion will use dimension-
√
less quantities. Lengths, body forces, spin rates and stress
πρG and πρ2GR2,
tensors are normalized by R, πρGR,
respectively. The normalized spin rate and radial distance
from the center are denoted as ω and Rb, respectively.
The spherical body is assumed to be rotating along a
constant spin axis. Here, this spin axis is lined up in the z
axis. The x axis is then given in an arbitrary direction on the
equatorial plane, and the y axis is defined to satisfy the or-
thogonal relation with these axes. This Cartesian coordinate
frame is later denoted as (x, y, z). We also use a spherical
coordinate system (Rb, θ, ψ), where θ is the latitude and ψ
is the longitude.
Material behavior is assumed to be elastic-perfectly
plastic -- that is, linearly elastic below the yield condition
and no-material hardening on it. To describe such behavior,
we use three material properties: Poisson's ratio ν, a fric-
tion angle φ and cohesive strength Y . Although we evaluate
a stress field based on linear elasticity, since our focus is only
on the stress, Young's modulus does not appear in our for-
mulation. Thus, we do not specify this value here. We also
consider the latitude θ a free parameter in the theoretical
investigation.
In this elastic-perfectly plastic model, only elastic strain
contributes to stress (Simo & Hughes 1998). Also, at the
yield condition stress does not change, and thus only plastic
strain occurs (Chen & Han 1988). On the assumption that
plastic deformation is small, this fact guarantees that the
yield condition of an elastic stress predicts when and where
plastic deformation appears and propagates. To calculate
an elastic solution for a spherical body, again, we use the
analytical form derived by Dobrovolskis (1982).
The present study needs to take into account the fact
that plastic deformation is a function of loading paths, which
are time-dependent. However, it takes about million years
for the YORP effect substantially to change the spin rate
of an asteroid (Rubincam 2000), implying that the changes
due to time are negligible. A recent breakthrough about the
stochastic YORP effect may also provide an interpretation
that an asteroid would rather change its spin state stochas-
tically than spin up constantly (Statler 2009; Bottke et al.
2015). This intimates that the internal structure may go
back and forth between elastic and plastic states.
MNRAS 000, 1 -- 10 (2002)
Holsapple (2010) who analyzed plastic deformation of
a body due to the YORP effect simplified his analysis with
the following crucial statement, "Clearly the possible anal-
yses are vast in parameter space, and are much too compli-
cated to expect closed-form solutions. A much simpler and
reasonable approach is to ignore the details of the applica-
tion and timing of those surface forces, and study the case
where the angular momentum is slowly increased in some
unspecified but quasi-static way." We follow his assumption
to consider the present problem a quasi-static problem. We
leave analyses about dynamical evolution of the stress field
in an asteroid as our future work.
We define the terminology about cohesive strength used
in this study. The critical cohesive strength is a computed
threshold that can prevent an asteroid from experiencing
failure. The "actual" cohesive strength, on the other hand,
is an assumed quantity that an asteroid may actually have,
which will be constrained by the critical cohesive strength.
Also, the spin rate, ω = (cid:112)4/3 ∼ 1.15, will be used to de-
scribe the condition at which a small particle resting on the
surface of a sphere has zero net acceleration (Hirabayashi
et al. 2015). This simply implies that beyond this spin rate,
materials there experience tension and are shed.
3 ANALYTICAL INVESTIGATION
3.1 Yield condition
Although Dobrovolskis (1982) defined duress to express the
dependence of materials on pressure and shear, here, we use
the Drucker-Prager yield criterion, a smooth function that
takes account of shear and pressure:
f = αI1 +
√
J2 − s (cid:54) 0.
(1)
I1 and J2 are the stress invariants:
I1 = σ1 + σ2 + σ3,
(2)
{(σ1 − σ2)2 + (σ2 − σ3)2 + (σ3 − σ1)2}, (3)
J2 =
1
6
where σi (i = 1, 2, 3) is the principal stress component. α and
s, which are functions of φ and Y , characterize the shape of
this yield surface. In general, the radius on the deviatoric
stress plane is determined by choosing appropriate α and s.
To determine these parameters, Chen & Han (1988) com-
pared the radius of the Drucker-Prager criterion with that
of the Mohr-Coulomb yield criterion. Here, we use the fact
that the stress field in a spherical object is always located at
the compression meridian, i.e., σ1 = σ2 > σ3 (Hirabayashi
& Scheeres 2015). Then, α and s are (Chen & Han 1988):
α =
√
2 sin φ
3(3 − sin φ)
, s =
√
6Y cos φ
3(3 − sin φ)
.
(4)
Failure of spherical bodies due to rotation
3
field at ψ = 0:
σexx = k1(1 − R2
σeyy = k1(1 − R2
σezz = k9(1 − R2
σexz = k14R2
b cos θ sin θ.
b ) − k14R2
b ) − k13R2
b ) − k14R2
b sin2 θ,
b cos2 θ − k14R2
b cos2 θ,
(5)
b sin2 θ, (6)
(7)
where
ω2
5
k1 =
k9 = − ω2
k13 = − 2ω2
k14 = − ω2
5
5
5
15
(1−ν)(7+5ν) − 2
12+ν−5ν2
(1−ν)(7+5ν) − 2
3−6ν−5ν2
4−3ν−5ν2
(1−ν)(7+5ν) + 4
3−6ν−5ν2
(1−ν)(7+5ν) + 4
15
3−ν
1−ν ,
3−ν
1−ν ,
1−2ν
1−ν ,
1−2ν
1−ν .
15
15
(8)
(9)
(10)
(11)
(12)
The indices of ks are chosen to be consistent with those by
Hirabayashi et al. (2015).
3.3 Characterization of failure conditions
The yield condition of the stress field is investigated to char-
acterize when and where failure occurs in the body. Specifi-
cally, we consider the critical cohesive strength, denoted as
Y ∗. From the equal condition of Form (1), we obtain Y ∗ as
a function of ω, φ, θ, Rb and ν:
∗
Y
(ω, φ, θ, Rb, ν) =
√
3(3 − sin φ)
6 cos φ
√
J2).
(αI1 +
(13)
As discussed in the previous section, since we assume that a
material is elastic-perfectly plastic, if Y ∗ at a given location
is higher than the actual cohesive strength, that location
reaches the yield. Thus, giving the contour curves of Y ∗
predicts the failed regions in the body. A negative value
of Y ∗ indicates that the locations having it do not need
cohesive strength to avoid the yield.
We analytically explore how Y ∗ evolves as ω changes.
To do so we compute Y ∗s at the body's origin and on the
surface, later denoted as Y ∗
s , respectively. At the
origin, where Rb = 0, the stress invariants are:
c and Y ∗
Then, we obtain Y ∗
c :
∗
c (ω, φ, ν) =
Y
I1 = 2k1 + k9,
J2 =
1
3
(k1 − k9)2.
(cid:20) 3 + 2ν
7 + 5ν
ω2
(3 − ν)(−2 + ω2)
1 − ν
3 − sin φ
6 cos φ
√
+
3
5
α
(14)
(15)
(cid:21)
.
(16)
Needless to say, Y ∗
spin rate should be:
c is independent of θ. When Y ∗
c = 0, the
3.2 Elastic stress solution in a spherical body
∗2
c =
ω
1 + 5√
3α
2
1−ν
3−ν
.
3+2ν
7+5ν
(17)
Hirabayashi et al. (2015) used a technique by Dobrovolskis
(1982) to provide an elastic stress in a spherical body rotat-
ing at spin rate ω. Here, we directly apply their formulation
to the present problem. Since the body forces acting on the
body are symmetric along the z axis, we consider the stress
This condition is a key parameter that tells us when the
central region needs cohesive strength to avoid failure. This
parameter indicates a lower threshold showing when a spher-
ical body becomes oblate.
On the surface, where Rb = 1, on the other hand, the
MNRAS 000, 1 -- 10 (2002)
4
Hirabayashi
stress invariants are:
I1 = −(k13 + k14) cos2 θ − 2k14 sin2 θ,
14) cos4 θ + k2
13 − k13k14 + k2
{(k2
1
3
+k14(k13 + k13) sin2 θ cos2 θ}.
J2 =
(18)
14 sin4 θ
s and the spin rate at which Y ∗
(19)
Then, Y ∗
s = 0, denoted as ω∗
s ,
are obtained as functions of θ by substituting these equa-
tions into Equation (13); however, since the expression is too
long to write down here, we only mention that the derivation
is trivial.
3.4 Results
c
c = Y ∗
c increases. Thus, ω∗
Figure 1 shows the contour plots of Y ∗
Hereafter, following the fact that the failure conditions are
not strong functions of Poisson's ratio (Holsapple 2008;
Hirabayashi & Scheeres 2014), we fix Poisson's ratio at 0.25,
which allows for compressibility of the volume. For φ, we
consider a range of 25◦ − 45◦ for a typical geological mate-
rial (Lambe & Whitman 1969).
(the dashed
s (the solid curves) in a θ − ω space. Figures
curves) and Y ∗
1(a), 1(b) and 1(c) display the cases of φ = 25◦, 35◦ and 45◦,
respectively. The red solid and dashed curves provide ω∗
s
and ω∗
c , respectively. The blue curves show the condition at
which Y ∗
s , which means that if the spin rate is higher
(lower), the central region is more (less) sensitive to failure
than the surface. The fact that the blue line and ω∗
c are close
to one another shows how rapidly Y ∗
c can
be considered to a parameter describing when the central
region starts to fail structurally. For the cases of φ = 25◦
and 35◦, Y ∗
s is always positive. The analytical expression of
s shows that it is always positive if 1 − 12α2 > 0, i.e.,
Y ∗
φ < sin−1(3/5) ∼ 36.87◦.
We plot the distributions of Y ∗ on the x − z plane for
different spin rates and friction angles in Figure 2, which are
analogous to Figure 4 in Dobrovolskis (1982) except that he
displayed the maximum shear. The left and right columns
indicate spin rates of 0.9 and 1.15, respectively. The 0.9 spin
rate is chosen based on ω∗
c . Figures 2(a) and 2(b), 2(c) and
2(d), and 2(e) and 2(f) display the cases of friction angles of
25◦, 35◦, and 45◦, respectively.
c = 0.85 and
Y ∗
s > 0 everywhere. At a spin rate of 0.9, if the actual cohe-
sive strength, later simply denoted as Y , is less than 0.02, a
region enclosing the surface and the central region becomes
the most sensitive to failure (Figure 2(a)). We note that if
the spin rate is less than ω∗
c , the central region should be
stable. At a spin rate of 1.15, however, since Y ∗ becomes
maximum at the central region, if Y is less than, for exam-
ple, 0.175, the central region where Y ∗ > 0.175 should fail
structurally (Figure 2(b)).
c = 0.95
and Y ∗
is positive everywhere, the blue curve is always
higher than, but close to, ω∗
c (Figure 1(b)). Contrast to the
25◦ friction angle case, at a spin rate of 0.9, since Y ∗ is nega-
tive at the central region, if Y is less 0.02, only the equatorial
surface should fail (Figure 2(c)). At a spin rate of 1.15, on
the other hand, and Y ∗ is the highest at the central region
(Figure 2(d)). For the case of a friction angle of 45◦, ω∗
c for
this case is 1.01. The results show the same features as the
For the case of a friction angle of 35◦, since ω∗
For the case of a friction angle of 25◦, ω∗
s
35◦ friction angle case; however, the region where Y ∗ > 0
only appears at the equatorial edges. ω∗
c s for these cases are
always lower than ω ∼ 1.15.
These cases describe that at high spin rates, a spher-
ical body requires higher cohesive strength to avoid fail-
ure of the central region. This corresponds to earlier stud-
ies (Hirabayashi & Scheeres 2015; Hirabayashi et al. 2015).
Plastic deformation should be compressive along the spin
axis and tensile on the equatorial plane. However, if the spin
rate is lower than ω∗
c , the equatorial surface is the most sen-
sitive, which is one of the findings in this study. A variation
of the friction angle causes a change of Y ∗; the higher the
friction angle is, the smaller Y ∗ is. As a result, ω∗
c increases
because a higher friction angle can prevent the body from
failing at faster spin rates. For any cases, however, Y ∗ of
the equatorial surface (θ = 0◦) becomes positive at slower
spin rates than that of the central region; the failure regions
move from the equatorial surface to the central region as the
spin rate increases (Figure 1). This result indicates that the
failure mode is not a strong function of a friction angle. It
should be noted that this results from linear elasticity, which
does not take into account plastic flow yet. Since plastic flow
can make wider regions fail, the actual failure region should
be wider than the current prediction (Chen & Han 1988).
More precise propagation of plastic deformation in the body
has to be discussed based on the flow rules, and we leave this
discussion as our future work.
4 NUMERICAL INVESTIGATION BY
PLASTIC FEM
We use a plastic FEM developed by Hirabayashi & Scheeres
(2015) to evaluate the analytical results above. The code is
developed by using commercial finite element program AN-
SYS 15.03. In this model, elastic behavior is linear, while
plastic behavior is perfectly plastic and follows an associ-
ated flow rule, which defines that plastic deformation al-
ways occurs in the normal direction to the yield surface. It
is also assumed that plastic deformation is small. We cre-
ate a 10-node tetrahedral FEM mesh for a spherical body.
Six node displacements are fixed so that the rotational and
translational motions are cancelled out. Similar to the ana-
lytical model, we consider materials in the body to be ho-
mogenous. The yield condition is the Drucker-Prager yield
criterion, which is described in Equations (1) through (4).
Figure 3 shows the results by the plastic FEM for the
same cases as Figure 2. The contours indicate the stress
ratio, a ratio of the current stress to the yield stress (Kohnke
2009); if this ratio is 1, the region should experience plastic
deformation, which is given in red. To induce the plastic
deformation seen in this figure we choose the actual cohesive
strength for each case.
For the case of φ = 25◦, we choose Y = 0.02 at ω =
0.9 to induce failure in both the central and surface regions
(Figure 3(a)) and Y = 0.175 at ω = 1.15 to produce that in
the central region (Figure 3(b)). For the case of φ = 35◦, we
input Y = 0.02 at ω = 0.9 to trigger failure on the equatorial
surface (Figure 2(c)), and Y = 0.15 at ω = 1.15 to make
the central region fail (Figure 2(d)). Last, for the case of
φ = 45◦, Y = 0.02 at ω = 0.9, and Y = 0.125 at ω = 1.15.
The case for ω = 0.9 does not have any failed regions in
MNRAS 000, 1 -- 10 (2002)
Failure of spherical bodies due to rotation
5
(a)
(b)
(c)
Figure 1. Contour curves of Y ∗. The solid and dashed curves indicate Y ∗
c , respectively. Figures 1(a), 1(b) and 1(c) describe
the cases of φ = 25◦, 35◦ and 45◦, respectively. The red curves indicate the conditions of zero-cohesive strength for both cases. The blue
lines give ω at which Y ∗
s . The color version is available online.
c = Y ∗
s and Y ∗
the body, while that for ω = 1.15 causes the similar failure
mode appearing in the central region. We confirm that the
sizes of the failure regions by the plastic FEM are always
wider than those by the analytical study. Regardless of this,
all the results by both of the techniques are fairly consistent.
We have to note that Figure 3 shows the deformed shapes
after loading.
5 DISRUPTION SCENARIO OF A
SPHERICAL BODY
The present analysis provided insights into evolution of a co-
hesive, spherical body due to the YORP spin-up. Using the
results obtained above, this section discusses possible evolu-
tion and disruption scenarios of a spherical body by taking
into account a two-layered-cohesive model by Hirabayashi
et al. (2015). On the assumption that the bulk density is
MNRAS 000, 1 -- 10 (2002)
6
Hirabayashi
Figure 2. Distribution of Y ∗ across the quadrantal cross section along the x and z axes. The elastic solutions are used to provide this
distribution. The left and right columns describe spin rates of 0.9 and 1.15, respectively. The first, second and third rows indicate friction
angles of 25◦, 35◦ and 45◦, respectively. The color version is available online.
homogeneous, their model consists of a spherical core which
has higher cohesive strength and a surface shell which pos-
sesses low cohesive strength. They concluded that if there is
a strong internal core, a possible disruption mode is surface
shedding.
Figure 4 indicates evolution paths of a spherical body.
Along every path (from the left to the right), the actual
c and ω ∼ 1.15 are
cohesion is considered to be constant. ω∗
used to characterize the stages of the evolution. A spherical
body with homogeneous structure would have two pathways
to its terminal state. If the actual cohesive strength is high
enough for the body to support its structure at a slow spin
rate, there are no failed regions in the body (A1). As the
spin rate becomes higher than ω∗
c , the central region then
reaches the yield first (A2). At this spin rate, although the
failed region becomes wider, since the body can still sustain
MNRAS 000, 1 -- 10 (2002)
Failure of spherical bodies due to rotation
7
Figure 3. Failure regions on the cross section. The horizontal and vertical axes are identical to the x and z axes, respectively. The plastic
FEM simulations are conduced to obtain these failure regions. The simulation conditions displayed here are the same as Figure 2. The
color version is available online.
itself, disruption does not occur yet. If the actual cohesive
strength is not high enough, stage A1 can result in stage
B2, which is discussed later. The body at stage A2 reaches
its terminal state at a spin rate higher than ω ∼ 1.15 (A3).
The actual cohesive strength is so high that the body can
spin fast; however, once the structure fails structurally at
a high spin rate, large components immediately fly away.
This disruption mode is a breakup of the body into multiple
components (Hirabayashi et al. 2015).
If the actual cohesive strength is low at a low spin rate,
the equatorial surface fails first (B1). Stage B2 is a phase
in which both surface and central regions fail at a spin rate
higher than ω∗
c , leading to propagation of plastic deforma-
tion over the entire equatorial plane. This stage can come
from two stages, A1 and B1. At this spin rate, since the equa-
torial plane has already failed, the body becomes oblate due
to vertical compression (Hirabayashi & Scheeres 2015). Once
the spin rate is higher than ω ∼ 1.15, the shape is more and
more oblate (B3). The oblateness may depend on a friction
angle because higher friction can prevent local elements from
deforming inelastically (personal communication with Paul
S´anchez, 2015). However, since plastic deformation highly
depends on loading paths (Chen & Han 1988), if the body
stays at critical conditions for a long time, the oblateness
may grow critically even for high friction angle cases. Since
the surface has failed already, a large amount of mass may
MNRAS 000, 1 -- 10 (2002)
8
Hirabayashi
be ejected from the edges (S´anchez & Scheeres 2012). The
prediction of this mode can be seen from Figures 2 and 3; if
the actual cohesive strength is small, failure widely spreads
out across the surface and central regions, leading to ver-
tical compression (Hirabayashi & Scheeres 2015) and mass
ejection.
If the body has a strong core and weak surface, it has
a different pathway to its terminal state (Hirabayashi et al.
2015). At a spin rate of less than ω∗
c , a failed region appears
on the equatorial surface first (C1). This mode should be
identical to stage B1. However, as the spin period becomes
higher than ω∗
c , since the central region is strong enough,
the surface region suffers failure (C2). A considerable fail-
ure mode at this stage is a landslide because the central
region has high cohesive strength to resist failure, and the
surface region should fail due to its low cohesive strength.
The shape at stage C2 may look similar to that at stage B2.
However, since the formation process is different, it may be
possible to distinguish between stages B2 and C2 by observ-
ing different surface morphology. The shape at stage B2 may
not have the morphology created by landslides, while that
at stage C2 may have. Once the spin rate is above ω ∼ 1.15,
only the surface mass is shed without large deformation of
the original body (C3). This disruption can be distinguished
from stage B3 because the amount of ejected mass and the
final shape can be controlled by the size of the internal core.
If the core is large, inelastic deformation is limited and only
a limited amount of mass can be ejected from the surface.
6 DISCUSSION
The present analysis shows that the disruption modes of as-
teroids may be dependent on the internal structure. On the
assumption that reported active asteroids are spherical, the
evolution diagram in Figure 4 shows that an asteroid hav-
ing a breakup, such as P/2013 R3, should be structurally
strong enough to fail at a high spin rate (A3). 62412 and
P/2012 F5 (GIBBS) may be between stages B3 and C3 be-
cause large components were not seen in their dust tails
(Sheppard & Trujillo 2015; Drahus et al. 2015). Again, we
emphasize that this is given only under the assumption that
they are spherical at this point. In fact, they would be elon-
gated rather than spherical. Further investigation is neces-
sary to explore elongated cases. Another interesting aster-
oid is 311P/PANSTARRS, formerly P/2013 P5, which has
ejected a small amount of mass from the body (Jewitt et al.
2013, 2015). This mode is consistent with stage C3. Also,
the size of dust particles was only a few hundred microns.
This gives rise to a possible scenario that the spin period of
this body is not too fast, so ejected boulders may not have
enough kinetic energy to escape from the main body. Note
that our study does not explain the episodic event of this
object. Again, further investigation is necessary to explain
this unique event.
It is interesting to consider possible formation scenar-
ios for a quasi-spherical object with equatorial ridges. Either
stage B2 or stage C2 may represent the observed shapes such
as 1999 KW4 Alpha (Ostro et al. 2006), 1950 DA (Busch
et al. 2007) and 2008 EV5 (Busch et al. 2011). To have
these stages, they should go through one of the following
processes: A1 → B2, B1 → B2 and C1 → C2. To go through
process B1 → B2, a homogeneous body has to have small
cohesive strength. Based on Figure 2(c), for a friction an-
gle of 35◦, a spherical body should have an actual cohesive
strength of less than 0.02 at a spin rate of 0.9. For these
asteroids, we find that the actual cohesive strengths of 1999
KW4 Alpha, 1950 DA and 2008 EV5 should be less than
8 Pa, 5 Pa and 0.5 Pa, respectively. For 2008 EV5, we as-
sume its bulk density to be 2000 kg/m3. Note that these
values are quite small compared to the reported cohesive
strength of 1950 DA, which is > 75 − 85 Pa (Hirabayashi &
Scheeres 2015), and that of P/2013 R3, which is 40 − 210
Pa (Hirabayashi et al. 2014).
We also investigate ω∗
c for these asteroids. The central
region begins to deform at this spin state and becomes more
oblate as the spin rate increases. Assuming a friction angle
to be 35◦, we obtain this boundary as ∼ 3 hours for 1999
KW4 Alpha, 1950 DA and 2008 EV5. Since the spin periods
of 1999 KW4 Alpha and 1950 DA are 2.7 hours and 2.1
hours (Ostro et al. 2006; Busch et al. 2007), respectively,
these asteroids can become more oblate due to failure of
their central regions at the current spin states. On the other
hand, the spin state of 2008 EV5, which is 3.8 hours (Busch
et al. 2011), is below this lower threshold, implying that the
body may structurally relax. However, because of its shape,
it might have been spinning at a spin period of shorter than
3 hours.
Finally, we mention the earlier studies about the shape
evolution of a cohesionless, spherical body. Harris et al.
(2009) developed a numerical model that computes an ideal
shape of a spherical body by considering the angle of repose.
Extending their study analytically, Scheeres (2015) found
that in the two-dimensional model, once failure occurs on
the surface, the material that comes off of the mid-latitudes
should leave behind material at the angle of repose for the
regolith. This is different from our results that showed that
the equatorial surface should fail first at a slow spin rate.
The difference between our work and his may come from
our computational settings. Specifically, our boundary con-
dition may be different from his. Our surface condition is
based on a zero-traction condition (Hirabayashi et al. 2015),
while he defined the surface stress based on the force bal-
ance on the surface. These are not necessarily the same. We
will investigate the connection between our work and his in
the future.
7 CONCLUSION
This paper explored how failed regions in a spherical body
evolve as the spin rate changes and deduced its termi-
nal state based on our results and the earlier works by
Hirabayashi & Scheeres (2015) and by Hirabayashi et al.
(2015). The main contributions are as follows:
consistent with the results by a plastic FEM.
• Prediction of the failed regions by the elastic model is
• The surface region at the equator needs higher cohesive
strength to avoid failure at a low spin rate, while the central
region must have much higher cohesive strength as the spin
rate increases.
• There is a spin rate condition, given as ω∗
c , at which the
central region needs cohesive strength to avoid failure. If the
MNRAS 000, 1 -- 10 (2002)
Failure of spherical bodies due to rotation
9
Figure 4. Evolution and disruption scenarios of a spherical body. The blue and yellow markers show elastic components, and the orange
markers describe failed ones. The blue components are structurally weaker than the yellow components. As the spin rate increases, the
disruption mode evolves differently due to the internal structure. The color version is available online.
spin rate is higher than this condition, a spherical body may
become oblate.
• ω∗
c increases as the friction angle becomes higher; the
friction angle changes the failure conditions. However, even
at different friction angles, the failed regions consistently
move from the equatorial surface to the central region as the
spin rate increases; the failure modes are not strong func-
tions of it.
• Three possible disruptions were discussed in this study.
For a homogeneous body, a strong structure allows the body
to remain at a higher spin rate and then causes it to have a
breakup into multiple components. A weak structure leads
to failure on the equatorial surface at a slow spin rate, and
then has failure propagation to the central region. As a re-
sult, at its terminal state, the body becomes oblate. For a
body with a strong core, a possible failure mode is only sur-
face shedding.
• The formation of the equatorial ridges on a spherical
body depends on cohesive strength. If the surface region
is very weak, the ridge should be made only by downslope
movement, which mainly occurs at lower latitudes; other-
wise, the main driver for it is inelastic deformation along
the spin axis.
MNRAS 000, 1 -- 10 (2002)
ACKNOWLEDGEMENTS
The author wishes to thank Dr. Patrick Taylor and Dr. Paul
S´anchez for useful comments on this study and Dr. Tony Do-
brovolskis for his careful and constructive reviews that im-
proved this paper. The license of ANSYS 15.03 is owned by
Aerospace Engineering Sciences, The University of Colorado
at Boulder.
REFERENCES
Bottke W. F., Vokrouhlick`y D., Walsh K. J., Delbo M., Michel P.,
Lauretta D. S., Campins H., Connolly H. C., Scheeres D. J.,
Chelsey S. R., 2015, Icarus, 247, 191-217
Brozovi´c M., Benner L. A., Taylor P. A., Nolan M. C., Howell
E. S., Magri C., Scheeres D. J., Giorgini J. D., Pollock J. T.,
Pravec P., Gal´ad A., Fang J., Margot J.-L., Busch M. W.,
Shepard M. K., Reichart D. E., Ivarsen K. M., Haislip J. B.,
LaCluyze A. P., Jao J., Slade M. A., Lawrence K. J., Hicks
M. D., 2011, Icarus, 216, 241-256
Busch M. W., Giorgini J. D., Ostro S. J., Benner L. A., Jurgens
R. F., Rose R., Hicks M. D., Pravec P., Kusnirak P., Ireland
M. J., Scheeres D. J., Broschart S. B., Magri C., Nolan M. C.,
Hine A. A., Margot J.-L., 2007, Icarus, 190, 608-621
Busch M. W., Ostro S. J., Benner L. A., Brozovic M., Giorgini
J. D., Jao J. S., Scheeres D. J., Magri C., Nolan M. C., Howell
E. S., Taylor P. A., Margot J.-L., Brisken W., 2011, Icarus,
212, 649-660
Chen W. F., Han D. J., 1988, Plasticity for Structural Engineers.
Springer-Verlag
10
Hirabayashi
Cotto-Figueroa D., Statler T. S., Richardson D. C., Tanga P.,
2015, The Astrophysical Journal, 803, 25 (18pp)
Davidsson B. J., 2001, Icarus, 149, 375-383
Dobrovolskis A. R., 1982, Icarus, 52, 136-148
Drahus M., Waniak W., Tendulkar S., Agarwal J., Jewitt D.,
Sheppard S. S., 2015, The Astrophysical Journal Letters, 802,
L8 (6pp)
Harris A. W., Fahnestock E. G., Pravec P., 2009, Icarus, 199,
310-318
Hirabayashi M., S´anchez D. P., Scheeres D. J., 2015, The Astro-
physical Journal, 808, 63 (12pp)
Hirabayashi M., Scheeres D., 2014, The Astrophysical Journal,
780, 160 (11pp)
Hirabayashi M., Scheeres D. J., 2015, The Astrophysical Journal
Letters, 798, L8 (5pp)
Hirabayashi M., Scheeres D. J., S´anchez D. P., Gabriel T., 2014,
The Astrophysical Journal Letters, 789, L12 (5pp)
Holsapple K., 2001, Icarus, 154, 432-448
Holsapple K. A., 2004, Icarus, 172, 272-303
Holsapple K. A., 2008, International journal of Non-Linear Me-
chanics, 43, 733-742
Holsapple K. A., 2010, Icarus, 205, 430-442
Jacobson S. A., Scheeres D. J., 2011, Icarus, 214, 161-178
Jewitt D., Agarwal J., Li J., Weaver H., Mutchler M., Larson S.,
2014, The Astrophysical Journal Letters, 784, L8 (5pp)
Jewitt D., Agarwal J., Weaver H., Mutchler M., Larson S., 2013,
The Astrophysical Journal Letters, 778, L21 (4pp)
Jewitt D., Agarwal J., Weaver H., Mutchler M., Larson S., 2015,
The Astrophysical Journal, 798, 109 (12pp)
Kohnke P., 2009, Theory Reference for the Mechanical APDL
and Mechanical Applications. ANSYS, Inc., Southpointe 275
Technology Drive Canonsburg, PA 15317, 12.1 edn
Lambe T. W., Whitman R. V., 1969, Soil Mechanics. John Wiley
& Sons
Marzari F., Rossi A., Scheeres D., 2011, Icarus, 214, 622-631
Ostro S. J., Margot J.-L., Benner L. A. M., Giorgini J. D.,
Scheeres D. J., Fahnestock E. G., Broschart S. B., Bellerose
J., Nolan M. C., Magri C., Pravec P., Scheirich P., Rose R.,
Jurgens R. F., De Jong E. M., Suzuki S., 2006, Science, 314,
1276-1280
Pravec P., Harris A. W., Warner B. D., 2006, Near Earth Objects,
our Celestial Neighbors: Opportunity and Risk, Proceedings
IAU Symposium, 2, 167-176
Rubincam D. P., 2000, Icarus, 148, 2-11
S´anchez D. P., Scheeres D. J., 2012, Icarus, 218, 876-894
Scheeres D., 2007, Icarus, 188, 430-450
Scheeres D., 2015, Icarus, 247, 1-17
Sharma I., 2013, Icarus, 223, 367-382
Sheppard S. S., Trujillo C., 2015, The Astronomical Journal, 149,
44 (9pp)
Simo J. C., Hughes T. J., 1998, Computational
inelasticity.
Springer-Verlag New York, Inc.
Statler T. S., 2009, Icarus, 202, 502-513
Veras D., Jacobson S. A., Gansicke B. T., 2014, Monthly Notices
of the Royal Astronomical Society, 445, 2794-2799
Walsh K. J., Richardson D. C., Michel P., 2008, Nature, 454,
188-191
Walsh K. J., Richardson D. C., Michel P., 2012, Icarus, 220, 514-
529
This paper has been typeset from a TEX/LATEX file prepared by
the author.
MNRAS 000, 1 -- 10 (2002)
|
1705.08633 | 1 | 1705 | 2017-05-24T07:08:03 | Discovery of a new branch of the Taurid meteoroid stream as a real source of potentially hazardous bodies | [
"astro-ph.EP"
] | Taurid meteor shower produces prolonged but usually low activity every October and November. In some years, however, the activity is significantly enhanced. Previous studies based on long-term activity statistics concluded that the enhancement is caused by a swarm of meteoroids locked in 7:2 resonance with Jupiter. Here we present precise data on 144 Taurid fireballs observed by new digital cameras of the European Fireball Network in the enhanced activity year 2015. Orbits of 113 fireballs show common characteristics and form together a well defined orbital structure, which we call new branch. We found that this branch is characterized by longitudes of perihelia lying between 155.9-160o and latitudes of perihelia between 4.2-5.7o. Semimajor axes are between 2.23-2.28 AU and indeed overlap with the 7:2 resonance. Eccentricities are in wide range 0.80-0.90. The orbits form a concentric ring in the inner solar system. The masses of the observed meteoroids were in a wide range from 0.1 g to more than 1000 kg. We found that all meteoroids larger than 300 g were very fragile, while those smaller than 30 g were much more compact. Based on orbital characteristics, we argue that asteroids 2015 TX24 and 2005 UR, both of diameters 200-300 meters, are direct members of the new branch. It is therefore very likely that the new branch contains also numerous still not discovered objects of decameter or even larger size. Since asteroids of sizes of tens to hundreds meters pose a treat to the ground even if they are intrinsically weak, impact hazard increases significantly when the Earth encounters the Taurid new branch every few years. Further studies leading to better description of this real source of potentially hazardous objects, which can be large enough to cause significant regional or even continental damage on the Earth, are therefore extremely important. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. Spurny_Taurids
May 25, 2017
c(cid:13)ESO 2017
Discovery of a new branch of the Taurid meteoroid stream as a real
source of potentially hazardous bodies
P. Spurný1, J. Borovicka1, H. Mucke2, and J. Svoren3
1 Astronomical Institute of the Czech Academy of Sciences, CZ-25165 Ondrejov, Czech Republic
e-mail: [email protected]
2 Astronomisches Büro, 1230 Wien, Austria
3 Astronomical Institute of the Slovak Academy of Sciences, SK-05960 Tatranská Lomnica, Slovak Republic
7
1
0
2
y
a
M
4
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
3
3
6
8
0
.
5
0
7
1
:
v
i
X
r
a
Received March 14, 2017; accepted May 1, 2017
ABSTRACT
Taurid meteor shower produces prolonged but usually low activity every October and November. In some years, however, the activity
is significantly enhanced. Previous studies based on long-term activity statistics concluded that the enhancement is caused by a swarm
of meteoroids locked in 7:2 resonance with Jupiter. Here we present precise data on 144 Taurid fireballs observed by new digital
cameras of the European Fireball Network in the enhanced activity year 2015. Orbits of 113 fireballs show common characteristics
and form together a well defined orbital structure, which we call new branch and which was evidently responsible for the enhanced
activity. This new branch is part of Southern Taurids and was encountered by the Earth between October 25 and November 17. We
found that this branch is characterized by longitudes of perihelia lying between 155.9 – 160◦ and latitudes of perihelia between 4.2 –
5.7◦. Semimajor axes are between 2.23 – 2.28 AU and indeed overlap with the 7:2 resonance. Eccentricities are in wide range 0.80 –
0.90. The most eccentric orbits with lowest perihelion distances were encountered at the beginning of the activity period. The orbits
form a concentric ring in the inner solar system. The masses of the observed meteoroids were in a wide range from 0.1 gram to more
than 1000 kg. We found that all meteoroids larger than 300 grams were very fragile (type IIIB), while those smaller than 30 grams
were much more compact (mostly of type II and some of them even type I). Based on orbital characteristics, we argue that asteroids
2015 TX24 and 2005 UR, both of diameters 200 – 300 meters, are direct members of the new branch. It is therefore very likely
that the new branch contains also numerous still not discovered objects of decameter or even larger size. Since asteroids of sizes of
tens to hundreds meters pose a treat to the ground even if they are intrinsically weak, impact hazard increases significantly when the
Earth encounters the Taurid new branch every few years. Further studies leading to better description of this real source of potentially
hazardous objects, which can be large enough to cause significant regional or even continental damage on the Earth, are therefore
extremely important.
Key words. Meteors, Meteoroids – asteroids – comets: 2P/Encke – Earth
1. Introduction
The Taurid meteoroid stream is one of the most studied mete-
oroid streams. This stream produces at least four meteor showers
on Earth: the Northern and Southern Taurids, both active from
end of September until December; the Daytime ζ-Perseids, ac-
tive from end of May to the beginning of July; and the Daytime
β-Taurids, active in June and the first half of July (Jenniskens
2006). Other showers may be also related to the Taurid stream,
namely the Piscids in September, χ-Orionids in December, and
Daytime May Arietids in May (Jenniskens 2006). Since the work
of Whipple (1940), the short period comet 2P/Encke has been
considered the most probable parent body of the Taurid stream.
It was, nevertheless, proposed that 2P/Encke is just a fragment of
a much larger comet, which was disrupted 103–104 years ago and
formed the whole Taurid complex including a number of aster-
oids (Clube & Napier 1984; Napier 2010). As more and more as-
teroids were being discovered over time, the number of asteroids
proposed by various authors as members of the Taurid complex
increased (e.g., Asher et al. 1993; Babadzhanov 2001; Porubcan
et al. 2006; Babadzhanov et al. 2008; Olech et al. 2016). The
problem is that the Taurid stream is very extended and the low-
inclination short-period Taurid orbits are very common for near-
Earth asteroids. Many proposed associations can be therefore
just random coincidences. Indeed, the spectra of six large as-
teroids proposed as members of the Taurid complex showed that
five of them are inconsistent with a cometary origin (Popescu et
al. 2014).
The activity of Taurids is prolonged but usually of low level.
In some years, however, the activity is enhanced, especially in
terms of large numbers of bright meteors (fireballs). Asher &
Clube (1993) proposed that there is a resonant swarm of mete-
oroids trapped in the 7:2 resonance with Jupiter. The expected
extent of the swarm was ± 30 – 40◦ in mean anomaly. Asher &
Izumi (1998) showed that enhanced Taurid activity indeed oc-
curred in the years when the center of the swarm was less than
40◦ in mean anomaly from the Earth at the beginning of Novem-
ber (the date of Taurid maximum). Asher & Clube (1993) pre-
dicted that future encounters would occur in 1995, 1998, 2005,
and 2008. In 1995, enhanced Taurid activity was observed by
the European Fireball Network (EN) when the rate of registered
Taurid fireballs was noticeably higher than is usual for the EN
at that time of year (Spurný 1996). Apart from regular Southern
and Northern Taurids, five fireballs observed during the last week
of October 1995 had distinct but very similar orbits. The radiants
lay near the regular southern Taurid radiant, but the initial veloc-
ities were larger (Vg = 33.1±0.3 km s−1). As a result these orbits
Article number, page 1 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fig. 1. Stations of the fireball network located in Czech Republic, Slovakia and Austria where DAFO are placed (status November 2015)
had significantly larger semimajor axes (a = 2.52±0.08 AU), ec-
centricities (e = 0.905±0.004) and inclinations (i = 6.2◦±0.4◦),
and smaller perihelion distances (q = 0.241±0.009 AU) than the
regular Southern Taurid orbit. The existence of this well-defined
cluster of similar Taurid meteoroids very probably means that
the enhanced activity in 1995 was caused by a new relatively
compact subsystem of the Taurid complex close to the South-
ern Taurids. As mentioned in Jenniskens (2006), this observation
identifies, for the first time, a meteor outburst associated with the
Taurid shower, and by implication the Earth crossing a relatively
young dust trail. The 1995 enhanced activity was also confirmed
by visual observations (Dubietis & Arlt 2007), both in terms of
increased overall activity and increased percentage of fireballs
in the period from October 23 to November 15 (McBeath 1999).
Similarly Johannink & Miskotte (2006), also from visual obser-
vations, confirmed increased Taurid activity in 1998 and 2005
and suggested that the Southern Taurids are responsible for the
higher activity in resonance years.
Shiba (2016) analyzed Taurid video observations from 2007
to 2015, including swarm encounter years 2008, 2012, and 2015
(see the webpage of D. Asher1 for swarm encounter predictions).
He confirmed that the enhanced activity is exclusively due to
Southern Taurids. Shiba also studied the dependency of orbital
elements on time and found that not only the mean orbital pe-
riod of swarm meteoroids but also that of Northern Taurids cor-
respond to the 7:2 resonance with Jupiter. The eccentricity was
found to decrease and perihelion distance to increase with time.
The work of Shiba work is statistical in nature, involving
thousands of meteors but with large individual uncertainties.
Here we present precise data on 144 Taurid fireballs observed by
new digital cameras of the European Fireball Network in 2015.
The description of the observational system, examples of the
data, and demonstration of their precision are given in Sect. 2
and 3. In Sect. 4 we show that the enhanced activity in 2015 was
1 http://star.arm.ac.uk/ dja/taurid/swarmyears.html
Article number, page 2 of 24
caused by a well-defined branch of Taurid meteoroids. We con-
centrate our study on orbital elements and only briefly discuss
the physical properties of the meteoroids. In Sect. 5 we show that
several known asteroids also belong to the branch, which caused
the 2015 activity. The implications of our work are discussed in
Sect. 6.
2. Observational techniques and data acquisition
The data reported here were obtained by the European Fireball
Network (EN). The core of the network, located in the Czech Re-
public, has been modernized several times (Spurný et al. 2007).
But the last significant improvement has been realized during
the last three years when a completely new instrument, the high-
resolution digital autonomous fireball observatory (DAFO), was
developed and gradually installed on the stations of the fireball
network between November 2013 and September 2015. These
new all-sky digital cameras are working alongside the older
analog (using photographic films) autonomous all-sky cameras
(AFO) on the majority of Czech stations but this older system
based on AFOs is gradually being decommissioned. At the end
of 2015 the DAFOs were installed on 13 stations around the
Czech Republic (Šindelová and Kocelovice stations are com-
pletely new and were built in mid-2015). Apart from the Czech
territory, two DAFOs were installed on the already working sta-
tions in Slovakia and Austria, respectively. The first DAFO was
installed at the observatory of the Slovak Academy of Sciences
in Tatranská Lomnica, where one AFO also remains in full oper-
ation, and the second, installed at the Waldviertel Observatory in
Martinsberg (Austria), substituted the previous AFO system in
September 2015. This core of the EN as schematically shown in
Figure 1 also cooperates with other parts and systems located in
neighboring European countries but data used in this study are
solely acquired by the stations based on the DAFO (vast majority
of used records) and AFO cameras as described above.
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Fig. 2. Detailed views of the EN051115_231201 Taurid fireball recorded by DAFOs at the stations 126 Martinsberg, 102 Kunžak, 107 Kucharovice,
and 114 Cervená hora. All-sky images from these stations were used for the analysis.
The imaging part of the DAFO system is comprised of a full
frame Canon 6D digital camera and a Sigma fish-eye lens (8
mm f/3.5) equipped with an electronic LCD shutter for speed
determination. In standard regime 16 interruptions and 35 s long
exposure are used. To avoid possible loss of data during read-
ing time of the CMOS sensor, we use two identical imaging sets,
which work in alternation mode with 5 second overlap. The older
AFOs analog imaging part is comprised of a Zeiss Distagon fish-
eye lens (30 mm f/3.5). Large format panchromatic sheet films
(9 x 12 cm, Ilford FP4) are used. The diameter of the sky on
the image is 8 cm and usually one exposure is taken per night.
Mechanical shutter with 15 interruptions per second is used. The
sensitivity limit is −4 magnitude for AFO (about 2−3 mag lower
around the full Moon period) and −2 magnitude for DAFO (with
lower dependence on lunar phase). Apart from the imaging part,
each DAFO and AFO is equipped with an all-sky radiometer
with time resolution of 5000 samples per second and with sim-
ilar sensitivity limit (in the moonless nights) like the imaging
system but with much higher dynamic range. These radiometers
serve several purposes, such as the real-time detection of fire-
balls, their exact absolute timing (system time is continuously
corrected by the PPS pulse of the GPS), recording of detailed
light curve profiles, and for precise photometry, especially for
brighter events when digital images become saturated as shown
in one example later.
The data presented in this study were obtained almost com-
pletely by the new digital autonomous system (DAFO). Thanks
to their higher sensitivity, fireball observations from DAFO con-
tain more information especially in the beginning and terminal
parts of the luminous trajectory in comparison with AFO. An-
other important advantage of DAFO is the ability to work dur-
ing periods when it is not completely dark (twilight periods) and
not completely clear (partly cloudy sky) as well. The data from
the new digital system allow us to reliably determine all basic
parameters of sufficiently bright fireballs up to the distance of
300 km from the stations (for special cases even up to 600 km).
It means that with the current number and displacement of sta-
tions (see Figure 1) we effectively cover territory of roughly 3
million square kilometers at least, i.e., a large part of Central Eu-
rope. All the advantages mentioned above significantly increased
the efficiency of our observations; and in direct comparison with
the efficiency of the previous analog AFO system the number of
recorded fireballs increased at least three times. When we com-
bine this increased efficiency with improved analysis techniques,
which we developed and gradually improved especially in the
last several years, we obtain results than were not reached by
any previous observing system used within the EN.
3. Data reduction
As described above, our fully automated instruments DAFO and
AFO provide us with two kinds of data: all-sky photographic
records and high-resolution radiometric light curves. For the
complete analysis of every fireball that was recorded from at
least two stations (the vast majority of the presented fireballs
were recorded from more than two stations) we use our own pro-
cedures, methods and analysis software. All-sky images taken
in the raw format are measured by the FishScan application,
which allows semiautomatic measurement of positions, speed,
and photometry. Usually the photometry from digital images is
reliable up to −8 apparent magnitude, brighter events start to be
saturated after reaching this brightness. However, thanks to the
Article number, page 3 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fig. 3. Lateral deviations of all measured points on the fireball luminous
path from the available records. The Y-axis scale is highly enlarged and
one standard deviation for any point on the fireball trajectory is only 7
m.
Fig. 4. Radiometric light curves of the EN051115_231201 fireball taken
by fast photometers (5000 samples/s) at Kunžak (blue) and Cervená
hora (red) stations. These apparent (not corrected for distance) light
curves taken from places185 km apart demonstrate perfect compliance
of both records; small differences in heights of individual peaks are
caused by different distances to the fireball.
high dynamic range of radiometers, which are incorporated in
each DAFO and AFO, we are able to obtain precise photometry
also for much brighter fireballs, even for superbolides as will be
shown later. We can calibrate radiometric records using not satu-
rated parts of the light curve obtained from photographic records.
Our whole procedure is demonstrated on the example below.
3.1. EN051115_231201 Taurid fireball: Example of data
analysis
The Taurid fireball of November 5, 2015, 23:12:01 UT, was
recorded photographically and photoelectrically at seven stations
in our network. For a complete analysis of this fireball, we chose
records taken from four stations that were close to its atmo-
spheric trajectory and were sufficient for reliable determination
of all parameters describing atmospheric trajectory, dynamics,
photometry, and heliocentric orbit of this fireball. A selection
of all-sky images of the fireball taken at individual stations are
shown in Figure 2.
The first step after measurement of all four digital images
and their astrometric reduction is computation of the atmo-
spheric luminous trajectory. We use two different methods de-
scribed in Ceplecha (1987): the so-called plane method, and in
Article number, page 4 of 24
5. Photographic
Fig.
light
EN051115_231201 fireball in absolute magnitudes.
and
radiometric
curves
of
the
Borovicka (1990), the so-called least-squares method. A first in-
dependent check of the results is that the values describing the
atmospheric trajectory obtained from these two methods turn out
to be the same within the uncertainties. Lateral deviations of all
measured points from the resulting atmospheric trajectory (zero
line) are shown in Figure 3. This plot illustrates the high reliabil-
ity of the astrometric solution. The spread of the measured points
from individual stations is random and the standard deviation is
only 7 m. In this context it is also important to mention how far
each station (camera) was from the fireball. Exact distances of
the beginning and terminal points R(B÷E) for each station were as
follows:
R(B÷E) (107) = 141.8 ÷ 92.7 km;
R(B÷E) (126) = 149.1 ÷ 119.5 km;
R(B÷E) (102) = 176.6 ÷ 143.3 km;
R(B÷E) (114) = 221.5 ÷ 189.8 km.
This example nicely illustrates that our records and meth-
ods provide us with a precision of the atmospheric trajectory de-
termination of about 10 m for fireballs that are still about 200
km away the stations. Most of the Taurids in this study, espe-
cially the fainter ones, were below or around this distance, and
only several of the brightest cases were at much larger distances
from the stations. The most distant Taurid was the superbolide
EN311015_180520, which was also recorded by the cameras at
distances of up to 630 km (stations that were used for analysis).
Data precision for such a distant and difficult to measure case is
about 140 meters, which is still good. The precision is crucial
not only for the determination of the position of the trajectory
in the atmosphere, but also for the determination of the direc-
tion of flight of the meteoroid, in other words, the position of the
apparent radiant, which is important for determination of the he-
liocentric orbit of the meteoroid. The coordinates of the apparent
radiant for this particular fireball were
αapp = 54.947◦ ± 0.007◦, δapp = 16.196◦ ± 0.004◦.
When computing the local azimuth and slope of the trajec-
tory, we took into account the curvature of bolide trajectory due
to gravity, which can be significant for longer fireballs with very
precise data. For the EN051115_231201 it is only 0.02◦.
The second step is the determination of the velocity of the
fireball. The data in this study are so good that it enabled us
to use the method described in Ceplecha et al. (1993) for the
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
vast majority of fireballs. Successful application of this model
is very sensitive to the quality of the lengths for each individ-
ual measured velocity point corresponding to a single shutter
break. This rigorous physical model provides the speed at any
point on the trajectory but for the presented study, which is fo-
cused on the orbital analysis, the initial velocity is the most
important. For the sample fireball we obtained a four parame-
ter (non-fragmenting) solution including initial velocity for each
station and all solutions were very similar. Nevertheless for the
final dynamic solution and initial velocity determination, not
only for this particular case but for all cases in this study, we
used a slightly different approach. We put all measured shutter
breaks from all used images together (timescales on all DAFOs
are correlated) and we applied the Ceplecha method on this uni-
fied data set. This approach significantly increases reliability of
the resulted dynamic solution. It is useful especially for shorter
fireballs such as Taurids because they are moderately fast mete-
oroids of cometary origin (i.e., relatively fragile) and the number
of measured breaks on one image can be limited. Therefore, ev-
ery independent measurement can be very useful in obtaining a
reliable value of the initial velocity. Moreover, for some fireballs
the non-fragmenting solution applied to the whole trajectory was
not adequate and we had to omit the terminal part of the fireball
to obtain a realistic value of the initial velocity. This was also
the case of the sample Taurid fireball EN051115_231201 as can
be seen for example in Figure 4 where several bright flares cor-
responding to fragmentation events are clearly visible. The re-
sulting value of initial velocity of the EN051115_231201 Taurid
fireball is 31.221 ± 0.037 km s−1.
The next step in the analysis of the available records is the ex-
act photometry of the fireball. We have two different data types,
those from photographic records and radiometers, from which
we can determine the brightness of the fireball and its initial
mass based on photometry. As mentioned above, we measure
digital images in 14-bit raw format. We found that this limited
dynamic range of the used CMOS sensor is sufficient for fire-
balls with apparent magnitude up to about −8. Above this limit
the measured signal starts to be saturated. As shown in Table 5,
which contains basic physical data of the presented Taurids, this
method can be reliably used for about 75% of all cases. The re-
maining 25% of presented cases are such bright fireballs that
their digital images are partly or even almost completely sat-
urated. For such fireballs we have different methods to describe
their brightness. One solution to this problem is the use of simul-
taneous photographic images taken by the AFO on the film. The
response of the film emulsion is logarithmic, which means that
the photographic film has much higher dynamic range; we use
Ilford FP4 panchromatic films with sensitivity 125 ASA. This is
a quite straightforward method and we used it in few cases, but
a still much more appropriate and accurate way is the use of the
light curves taken by the radiometers, which are in our cameras
and still have much higher dynamic range. The apparent (i.e.,
not corrected for distance) high-resolution (5000 samples/s) ra-
diometric light curves of the EN051115_231201 fireball taken
by radiometers at Kunžak (blue) and Cervená hora (red) stations
are shown in Figure 4. The close agreement between the dif-
ferent records, a testament to the high precision of the data, is
evident. As for other two closer stations to the fireball, the ra-
diometric light curves have exactly the same profile and could
be used for fireball photometry. However, we cannot use radio-
metric light curves directly because individual radiometers have
different sensitivity and are not calibrated to obtain absolute pho-
tometry. For the purpose of calibrating these records we combine
photometry from both methods. We measure the meteor signal
on the digital image and for calibration we use that part of the
photographic image, which is not saturated and at the same time
well above the noise of the measured signal from both the pho-
tographic records and corresponding radiometric records. This
is usually somewhere in the interval between −4 and −7 magni-
tude. However, we have to relate the timescale of the photograph
to the absolute timescale of the radiometer. For this purpose we
use time marks (breaks of double length) made by the electronic
shutter along the luminous path of the recorded fireball on the be-
ginning of each second. This defines the exact absolute time of
this measured point, which can be simply identified with the cor-
responding point on the radiometric light curve. Both radiome-
ter and electronic shutter are continuously corrected by the PPS
pulse of the GPS so the absolute timing of both records is given
with high precision.
The result of this procedure is illustrated on Figure 5 in
which photographic and radiometric light curves from all used
stations in absolute magnitudes are plotted. The first evident re-
sult is that, especially for shorter and faster fireballs containing
bright and short significant flares, the photographic photome-
try cannot correctly describe the shape of the light curve be-
cause of the low time resolution. The electronic shutter, on the
other hand, has a resolution of 16 interruptions per second with
the same length for the on/off state. This means that blind time
lasts exactly 0.03125 seconds and it is sometimes longer than
the duration of a flare or at least its brightest part. As can be
seen in Figure 4 and Figure 5, this is exactly the case of fireball
EN051115_231201, where most of the light is contained within
five distinct and quite short flares that are only partly recorded
(or even missing) on the image. Another aspect that is evident
in Figure 5 is the saturation of the photographic records. The
absolute photographic photometry profile (brightest parts of the
light curves) is, unlike the radiometric photometry, quite differ-
ent for individual stations. However it confirms the saturation
effect because stations, which are more distant from the fireball
give higher absolute maximum brightness. It means that these
records are not as saturated as the records from closer stations.
The last aspect, which is worth mentioning in connection with
the photometry shown in Figure 5, is that for correct calibration
of radiometric light curves it is much better to use the radiomet-
ric record and photographic image from the closest station where
the signal-to-noise ratio is the most favorable. This is valid es-
pecially for the fireballs as in the case described here, when the
increase of the brightness is very steep and the suitable (not sat-
urated) interval of magnitudes is very short.
A general conclusion is that the high-resolution radiometric
records are crucial for correct recovery of the photometry of all
brighter fireballs, especially those that contain distinct flares. We
note that the photometry based on radiometric light curves was
determined for about 90% of analyzed Taurids in this study.
As explained above, apart from the precise photometry of the
recorded fireballs, radiometric records provide us with a very ac-
curate absolute time of each event. This important parameter is,
along with the initial velocity and radiant position, necessary for
reliable determination of the heliocentric orbit of the observed
fireball. The orbits were computed by the method of Ceplecha
(1987).
To compute the photometric mass of the meteoroid, the ve-
locity dependence of the luminous efficiency was taken from
ReVelle & Ceplecha (2001). The mass dependence was ignored
by substituting 10 kg for the mass in their formula. Specifically,
the luminous efficiency for velocities above 25.4 km s−1 was as-
sumed to be directly proportional to the velocity, reaching 6.5%
at 30 km s−1. A factor of 1500 W for zero magnitude meteor
Article number, page 5 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fig. 6. Exposure coverage representing observing conditions in Central Europe during activity of Taurids in 2015. It is the fraction of real time
when all cameras in the network exposed to their total prescribed exposure time. The date corresponds to the evening date of the whole night.
Fig. 7. Activity of Taurids recorded by the DAFO cameras of the European Fireball Network in 2015. The number of presented Taurid fireballs in
the plot is 143 (total number is 144); S TAU from 28.11.2015 is out of range of the date axis. Date corresponds to the evening date of the whole
night.
(Ceplecha et al. 1998) was used to convert magnitudes into bolo-
metrically radiated energy.
The above-mentioned example clearly demonstrates high
precision and reliability of all parameters describing the atmo-
spheric trajectory, dynamics, photometry, and heliocentric orbit
not only for this particular case, but also for all Taurid fireballs
presented in Table 4 and Table 5. For this complex analysis of all
presented Taurid fireballs, i.e., the astrometric reduction of the
images, atmospheric trajectory computation, dynamic and pho-
tometric solutions, and finally orbital calculations, we used our
new software package BOLTRACK (J. Borovicka).
Although the autumn weather, especially in November, is no-
toriously cloudy in Central Europe, the year 2015 was not so bad.
There were several clear nights, especially in the beginning of
November, and only a few nights were completely cloudy prac-
tically at all stations. This situation is illustrated in Figure 6 in
which the ratio of real to prescribed exposure time for all stations
in the network altogether and for each individual night cover-
ing the Taurids activity in 2015 is shown. Since the DAFOs also
work when the sky is only partly clear, some fireballs were cap-
tured even in the nights when it was mostly cloudy and could
be still used for this study. On the other hand, some fireballs
were recorded only from one station or their records were of a
quality that is insufficient to merit scientific analysis. Such cases
were excluded from our study. As a result of relatively favorable
weather conditions and the capability of our network, we were
able to cover the whole period of the enhanced Taurid activity
from the last decade of October to mid-November as shown in
Figure 7. It is difficult to construct the activity profile for such
a long interval from our data because it is difficult to take into
Article number, page 6 of 24
account all observational effects and correctly eliminate them.
So Figure 7 does not represent the real activity profile, but only
the uncorrected distribution of selected Taurid fireballs during
the whole interval of activity. As described in the following sec-
tions, we identified three different groups of Taurids in our data
set. These are the regular Southern and Northern Taurids (desig-
nated as S TAU and N TAU, respectively) and a new branch of
Southern Taurids, which we designate S TAU (SB). As shown
later, this new branch was responsible for the enhanced activity.
From Figure 7 we can see that enhanced Taurid activity caused
by this new branch of Southern Taurids fireballs started on Oc-
tober 24, culminated around November 5 and terminated on the
night of November 16/17. The S TAU (SB) was not observed
after this date even though on November 18 and 21-23 were at
least partly clear nights with good observing conditions. From
Figure 7 we also see that the enhanced activity increased gradu-
ally with several days of very high activity at the turn of October
and November. This interval was strongly affected by the full
Moon period, so fainter fireballs were below the sensitivity limit
of the digital all-sky cameras and radiometers especially in the
end of October. Therefore the number of fireballs on these nights
may be underestimated. On the other hand, the relatively steep
decrease of activity after November 5 seems to be real. With re-
gard to regular S and N Taurids, they are quite uniformly spread
over the entire interval of observed activity.
4. The 2015 Taurid data
The total number of Taurid fireballs recorded photographically
by our instruments at least from two stations in 2015 was about
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Fig. 8. Detailed view of the two brightest Taurids far over Poland recorded by the AFO (analog camera) at station Polom.
200. This is much more than we recorded in any previous year.
The main reason for this is evidently the unusually high Taurid
activity, but it is also caused by much more efficient observa-
tional system and also by a quite long period of relatively good
weather. For this study we selected 144 Taurids with complete
information about heliocentric orbits of individual meteoroids
and their physical properties as well. Our data set is unique not
because of the total number of used meteors but the high preci-
sion of the data for each individual case, which was obtained by
with high-resolution cameras and radiometers and elaborated re-
duction methods. Meteor records were reduced one-by-one and
all steps in the measurement and computation process were un-
der careful human supervision.
Before going to the statistical analysis of the whole data set,
we describe in more detail some remarkable fireballs.
4.1. Exceptional cases
4.1.1. EN311015_180520 and EN311015_231301 - two
brightest Taurids
It is a well-known fact that Taurids are quite rich in bright fire-
balls. However, 2015 was also in this aspect exceptional and our
cameras recorded several very bright Taurids during the whole
period of activity. Altogether 24 Taurids were brighter than −10
absolute magnitude and 10 were similarly bright or even brighter
than the full Moon. Moreover, two of those, both observed in the
first half of the night of October 31, are really remarkable not
only in this data set, but in all Taurids that we have recorded
within the EN until now. Both are shown in Figure 8, where a
small part of the all-sky image is shown. This image was taken
from station Polom by the AFO, i.e., on sheet film, where one
exposure was taken per night. Both bolides were observed in a
similar (northern) direction and flew over northern and central
Poland, respectively. The brighter bolide, which reached a peak
absolute magnitude of −18.6 (on the left), occurred at 18:05:20
UT and the second, with peak magnitude of −15.8, occurred 5
hours 7 minutes and 41 seconds later at 23:13:01 UT. This is the
reason why their directions of flight differ, although both bolides
had practically the same radiant. The first was so bright that
it belonged to the superbolide category. This spectacular Tau-
rid bolide was caused by a meteoroid with initial mass more
than 1000 kg, i.e., a meter-sized object. Because of its enormous
brightness, clear skies over large parts of Central Europe, and
convenient time of its occurrence (it was an unusually nice Sat-
urday evening), thousands of eyewitnesses were fascinated by
this extraordinary natural event. We obtained more reports of
one bolide than ever before. Apart from plenty of visual obser-
vations, all DAFOs and AFOs in our network (at 15 stations)
recorded it, which was crucial for reliable description of this su-
perbolide. In addition to our own photographic and radiomet-
ric records, we used also two casual images. The first one is
a high-resolution digital image from Studénka, Czech Repub-
lic, which was obtained from amateur astronomer B. Pelc. The
second digital image was taken by G. Zieleniecki at Czernice
Borowe, Poland, and was freely available on the internet. Alto-
gether we used 13 most suitable photographic and 5 radiometric
records. The situation with the second, much smaller, meteoroid
was similar. It was also recorded by all our cameras at all 15
stations, and we obtained also 4 high-resolution casual digital
images from northern region of the Czech Republic. These im-
ages, which we also partially used, were taken by T. Chlíbec at
Klínovec, L. Sklenár from Kuncice and Labem, D. Šcerba from
Dolní Údolí, and L. Shrbený from Rícany; this record also in-
cludes a spectrum of the bolide. In this case we used the best 12
photographic and 4 radiometric records for final analysis.
Since these fireballs were exceptional, we modeled them
with our semiempirical fragmentation model (Borovicka et al.
2013). The model fits radiometric curves and deceleration. This
way we obtained more reliable initial masses of meteoroids
(1300 kg and 34 kg, respectively) and insight into their at-
mospheric fragmentation. Both meteoroids were effectively de-
stroyed high in the atmosphere under dynamic pressures < 0.05
MPa. In both cases a small fragment (< 1 kg) survived the ini-
tial destruction and fragmented further under pressures of ∼ 0.1
MPa. In comparison with other bright bolides (Borovicka et al.
2017), both Taurids were extremely fragile.
Simultaneously with our network, both bolides were also
recorded by the cameras of the Polish Fireball Network. These
data were analyzed independently and were published by Olech
et al. (2016). Because our data differ from their data, we provide
here our complete results and compare them to those reported
in Olech et al. (2016). Atmospheric trajectories are given in Ta-
ble 1, light curves in Fig. 9, and heliocentric orbits in Table 2.
When computing the local azimuth and slope of the trajectory,
we took into account both the curvature of the Earth and the
curvature of the bolide trajectory due to gravity, which was sig-
nificant for EN311015_180520 (change of direction of flight by
0.17◦ over the recorded length). Azimuths are measured from the
south clockwise. The apparent radiants given in Table 2 are valid
for the average points on the trajectories.
As shown later in this paper and that of Olech et al. (2016)
(in fact that paper is based only on these two bolides) data about
Article number, page 7 of 24
Table 1. Atmospheric trajectory data for the EN311015_180520 (left) and EN311015_231301 (right) bolides.
A&A proofs: manuscript no. Spurny_Taurids
Beginning
Terminal
Beginning
57.644 ± 0.030
114.724 ± 0.025
33.07 ± 0.03
18.46416 ± 0.00018
53.60723 ± 0.00050
18.574 ± 0.014
267.240 ± 0.013
Height (km)
Velocity (km s−1)
Longitude (◦ E)
Latitude (◦ N)
Slope (◦)
Azimuth (◦)
Time1(s)
Total length (km)
1Time zero corresponds to 18:05:18 UT for EN311015_180520 and 23:13:00 UT for EN311015_231301.
15.82901 ± 0.00020
53.50244 ± 0.00048
17.148 ± 0.034
265.120 ± 0.015
22 ± 2
−1.07
4.68
120.026 ± 0.030
32.56 ± 0.09
18.18064 ± 0.00026
52.13173 ± 0.00060
53.28 ± 0.04
350.78 ± 0.04
−0.07
Max.
bright.
80.8
33.07
16.927
53.553
17.71
266.00
2.22
186.3
Terminal
57.305 ± 0.016
30 ± 2
18.07087 ± 0.00014
52.54390 ± 0.00033
52.89 ± 0.10
350.69 ± 0.05
2.36
Max.
bright.
74.4
32.53
18.101
52.430
53.00
350.72
1.68
78.5
Table 2. Apparent and geocentric radiants and orbital elements
(J2000.0) for the EN311015_180520 (left) and EN311015_231301
(right) meteoroids. Time is given for the average point of the recorded
trajectory.
Time (UT)
αR (◦)
δR (◦)
v∞ (km s−1)
αG (◦)
δG (◦)
vG (km s−1)
vH (km s−1)
a (AU)
e
q (AU)
Q (AU)
ω (◦)
Ω (◦)
i (◦)
P (yr)
Perihelion
TPJup
EN311015_180520
18h05m20.0s ± 0.1s
50.126 ± 0.009
16.452 ± 0.016
33.068 ± 0.030
51.692 ± 0.010
14.592 ± 0.017
30.869 ± 0.032
37.32 ± 0.02
2.250 ± 0.009
0.8724 ± 0.0006
0.28715 ± 0.00032
4.212 ± 0.018
121.687 ± 0.022
37.791
5.707 ± 0.023
3.375 ± 0.020
2.952 ± 0.009
2012-07-26 ± 7 d
EN311015_231301
23h13m01.5s ± 0.1s
51.853 ± 0.022
15.66 ± 0.04
32.56 ± 0.09
51.445 ± 0.022
14.49 ± 0.04
30.59 ± 0.10
37.34 ± 0.06
2.258 ± 0.027
0.8689 ± 0.0020
0.2960 ± 0.0010
4.22 ± 0.05
120.62 ± 0.06
38.005
5.62 ± 0.06
3.39 ± 0.06
2012-07-20 ± 22 d
2.953 ± 0.028
light
curves
radiometric
9. Calibrated
Fig.
bolides
EN311015_180520 and EN311015_231301 (solid curves). For
EN311015_180520 data from two imaging cameras are also given
(crosses). After bolide maximum, most of radiometric signal was
produced by a stationary trail. Camera data contain only the bolide
moving further down. Time zero corresponds to 18:05:18 UT for
EN311015_180520 and 23:13:00 UT for EN311015_231301.
of
these big bolides are of great importance. Since it may not be
simple to distinguish which data set is correct, we carry out an
analysis of the differences. The positions of the recorded be-
ginning and end points of the bolide depend on the sensitivity
of the instrument and the observing conditions. Nevertheless,
when plotting the trajectory of the first bolide on the map, the
solution of Olech et al. (2016) is shifted about 1.8 km to the
north. As for the observed apparent radiant, there is a difference
of 0.65◦ in declination, i.e., 10 times their quoted uncertainty
(σ). For the second bolide, the larger difference is in right as-
cension (0.24◦, 4σ) and especially in entry velocity, which is
larger by 0.6 km s−1 (6σ) in Olech et al. (2016). Although our
data were obtained from large distances, our results are based
on large number of records (in both cases more than 10) and the
solutions for both bolides are very consistent. The photograph
from Czernice Borowe, when combined with our cameras, pro-
vides convergence angles in excess of 60◦ for the first bolide.
Czech cameras have mutual convergence angles up to 15◦. For
the second bolide the situation is even better, although the veloc-
ity was more difficult to measure. So there is no reason, why our
results should be different by more than the standard deviations
given in Table 1.
Article number, page 8 of 24
There is also at least a 3-5 seconds difference in the reported
time of appearance of the first fireball. Our radiometers are con-
tinuously corrected by PPS pulse of GPS and their timing preci-
sion is in millisecond range. A nice example how our radiome-
ters are synchronized is shown in Figure 4. Another discrepancy
is in the determination of the maximum absolute brightness for
both bolides. While Olech et al. (2016) determined the maxi-
mum absolute brightness −16.0± 0.4 mag for the first bolide, we
found that it reached −18.6 ± 0.2 mag. Our method described in
Section 3.1 relies on the linearity of radiometers even for strong
signals. Results from five independent radiometers were in per-
fect agreement. Similarly the brightness of the second bolide was
underestimated by Olech et al. (2016) by about 1 magnitude.
Regardless of these differences in the directly determined tra-
jectory parameters, we found another discrepancy in the calcu-
lation of orbital elements in Tables 3-5 of Olech et al. (2016).
We obtained significantly different results than those published
in Tables 5 and 6 of Olech et al. (2016) when we took their in-
put values of initial velocity, apparent radiant position (which
we recalculated from J2000.0 to the date that the bolides oc-
curred) time, and mean position for both bolides from their Ta-
bles 3 and 4 and used our program for orbital calculation; orbits
from this program were independently validated for example in
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Fig. 10. Detailed view on the very long Taurid fireball recorded by the DAFO (digital camera) at station Kunžak.
Clark & Wiegert (2011). We found the following differences for
the EN311015_180520 bolide (computed minus published): ∆a
= 0.0077 AU, ∆e = 0.0057, ∆ω = 1.3◦ (!), ∆i = 0.07◦, and ∆P
= 0.019 yr. For the EN311015_231301 bolide, differences are
∆a = 0.1287 AU (!), ∆e = 0.0087, ∆ω = 0.1◦, ∆i = 0.035◦, and
∆P = 0.20 yr. Some of these differences are really high, namely
1.3◦ , in argument of perihelion for the first bolide and especially
0.128 AU in semimajor axis for the second bolide. With the ra-
diant and velocity given by Olech et al. (2016) this bolide would
be far from the 7:2 resonance with Jupiter, nevertheless, their
published orbit puts it in the resonance.
4.1.2. EN061115_164758: An almost horizontal Taurid
On November 6, 2015 during dawn, just after the Taurid radi-
ant rose above the horizon, a relatively faint Taurid fireball of
−5.1 maximum absolute magnitude traveled over a large part
of sky and was observed by several stations in the SW part
of our network. The sky was not completely dark, especially
from the stations in western part of the network, which were
closest to the fireball trajectory. However, thanks to the higher
sensitivity of the digital cameras, this extremely long fireball
was nicely recorded on three stations, Kunžak, Martinsberg and
Kucharovice, which enabled us to describe this exceptional Tau-
rid accurately. Owing to its small slope which was 7.7◦ at the be-
ginning and during the flight decreased to only 5.6◦, the recorded
fireball trajectory was extremely long, i.e., exactly 258.7 km, and
its flight lasted 8.5 seconds. It is the longest Taurid fireball we
have ever recorded, both in duration and length. Thanks to a large
amount of data points, this Taurid has the best dynamic data and
the trajectory, i.e., also radiant, is also very precise. The initial
velocity of 31.285 km s−1 was determined with a precision of
±7 m s−1. As for the brightest Taurid described in Sect. 4.1.1,
we took into account the curvature of the trajectory of the bolide
due to gravity, which was significant for such a long and low
inclined fireball (change of direction of flight by 0.18◦ over the
recorded length). A detailed view of its luminous flight taken by
the DAFO at Kunžak station is shown in Figure 10. Additional
information about this fireball is given in Tables 4 and 5.
4.2. Radiants and orbits
In this section the radiants, velocities, and heliocentric orbits of
all 144 fireballs are evaluated. All elements in this paper are
given for equinox J2000.0. The data are presented in Table 4.
Figure 11 shows the dependency of geocentric radiant and ve-
locity on solar longitude (i.e., the longitude of the Sun at the
time of fireball observation). Thirteen fireballs were classified as
Northern Taurids. They can be easily recognized by their radi-
ant lying to the north of the ecliptic. All other fireballs belong
to Southern Taurids. Among them, a well-defined structure can
be recognized, where the radiant position and velocity are strict
linear functions of solar longitude. We call this structure a new
branch. Evidently, this branch was responsible for the enhanced
Taurid activity in 2015.
(1)
(2)
(3)
Regular Taurids also exhibit radiant motion but the spread of
individual radiants is much larger than for the new branch. For
the new branch, we found the following relationships:
αg = 46.99◦ + 0.554 · (λ(cid:12) − 210◦)
δg = 14.00◦ + 0.060 · (λ(cid:12) − 210◦)
vg = 32.90 − 0.293 · (λ(cid:12) − 210◦),
where αg and δg are the right ascension and declination, respec-
tively, of the geocentric radiant (J2000.0), vg is the geocentric
velocity in km s−1, and λ(cid:12) is solar longitude (J2000.0). Although
we defined the new branch on the basis of orbital elements rather
than radiants and velocities (see below), all fireballs of the new
branch had the radiant right ascension within 1.3◦ and declina-
tion within 0.7◦ from (1) & (2). The velocities were within 0.9
km s−1 from relationship (3). We point out that this spread is real.
The precision of most of our data, as demonstrated in Sect. 3.1,
is < 0.05◦ in the radiant position and 0.1 km s−1 in the velocity.
Fireballs from the branch were observed between solar lon-
gitudes 211◦ – 234◦ (October 25 – November 17). Taurids ob-
served before and after these dates belong to the background
population of Northern and Southern Taurid streams.
Longitude of perihelion, inclination, eccentricity, and peri-
helion distance as a function of solar longitude are plotted in
Fig. 12. Most notably, there is a concentration of orbits with lon-
gitude of perihelion, π (π = Ω + ω, where Ω is longitude of
ascending node and ω is argument of perihelion), at 158◦ ± 2◦
(Fig. 12a). There is only a weak correlation with solar longitude.
Similarly, there is a concentration of orbits with inclinations of
5.5◦ ± 1◦ (Fig. 12b). Regular Taurids show much larger spread,
145 – 175◦ in π and 2 – 7◦ in inclination.
Eccentricities and perihelion distances of the members of the
new branch are steep functions of solar longitude (Figs. 12c,d),
Article number, page 9 of 24
A&A proofs: manuscript no. Spurny_Taurids
All eccentricities lie within 0.012 from (4) and perihelia lie
within 0.027 AU from (5). Again, regular Taurids show much
larger scatter.
The new branch is best recognized in the plot of longitude of
perihelion, π, versus latitude of perihelion, β (sin β = sin ω sin i,
where i is inclination), presented in Fig. 13. We can state that
the new branch has π between 155.9 – 160◦ and β between 4.2
– 5.7◦. For regular Southern Taurids the observed spread in β is
2.5 – 6.5◦. Northern Taurids have negative β.
Semimajor axes are plotted in Fig. 14. For regular Taurids,
they lie between 1.9 and 2.4 AU. According to the model of
Asher & Clube (1993), the enhanced activity is caused by mete-
oroids trapped in the 7:2 resonance with Jupiter. The resonance is
located at 2.256 AU and extends from about 2.231 AU to 2.281
AU (Asher & Clube 1993). With two exceptions, the semima-
jor axes of all meteoroids with longitudes and latitudes of per-
ihelia within the above-defined limits fall in the 7:2 resonance.
Only two meteoroids had significantly lower semimajor axes,
2.15 – 2.16 AU. We consider them to be interlopers from the
background population of Southern Taurids, although the 15:4
resonance located at 2.155 AU might be at work here.
On the contrary, some Southern Taurids with perihelia out-
side the new branch limits were also in the 7:2 resonance. As
seen in Fig. 13, all of the Souther Taurids had an orientation of
perihelia relatively close to the new branch. Nevertheless, some
Northern Taurids were in the 7:2 resonance as well and they were
far from the new branch.
There is no correlation between semimajor axis and solar
longitude. The Tisserand parameter with respect to Jupiter in-
creases with solar longitude from 2.9 to 3.1 within the new
branch. This is due to the decreasing eccentricity. The often cited
boundary at TJup = 3 or 3.05 (e.g., Tancredi 2014) has no signif-
icance in this case.
According to the above definitions based on perihelion ori-
entation and semimajor axis, there are 13 Northern Taurids in
our data set, 18 regular Southern Taurids, and 113 members of
the new branch.
It is evident that the new branch represents an orbital struc-
ture that is much more compact than regular Taurids. Since the
activity of the new branch lasted almost one month, it cannot,
however, be a narrow filament. In order to visualize the new
branch, we plotted selected orbits covering the whole activity pe-
riod in Fig. 15. Unlike usual meteoroid streams, where the orbits
near perihelion largely overlap, here we see a concentric ring of
orbits near perihelion, which is more than 0.2 AU wide. As the
Earth moves around the Sun, it encounters first the orbits with
smaller perihelia and larger eccentricities. With increasing solar
longitude, orbits with progressively larger perihelia and smaller
eccentricities are encountered. Since all of the semimajor axes
are similar, eccentric orbits have larger aphelia than less eccen-
tric orbits and the orbits therefore intersect at about 3.6 AU.
4.3. Physical properties
The Taurids in our sample reached maximum absolute magni-
tude between −2 and −18.6. The photometric masses range from
0.1 gram to 1300 kg, i.e., there is a range of 7 orders of magni-
tude in mass. The mass distribution is given in Fig. 16, which
shows that the new branch has a higher proportion of massive
meteoroids. The data in Fig. 16 are biased because brighter me-
teors could be observed over large distances and under worse
conditions than faint meteors, nevertheless, the bias is the same
for all branches.
Fig. 11. Position of geocentric radiant and geocentric velocity as a
function of solar longitude for Taurids observed in 2015. Northern and
Southern Taurids are shown by different symbols and the Southern Tau-
rids belonging to the new branch are highlighted in light blue. The error
of the data is smaller than the size of the symbols in most cases.
i.e.,
e = 0.901 − 0.00403 · (λ(cid:12) − 210◦)
q = 0.224 + 0.0092 · (λ(cid:12) − 210◦).
Article number, page 10 of 24
(4)
(5)
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Fig. 12. Selected orbital elements as a function of solar longitude for Taurids observed in 2015. The symbols are the same as in Fig. 11. Errors are
in most cases smaller than symbol sizes and for clarity are plotted only for regular Taurids.
The beginning, maximum brightness, and end heights of all
studied fireballs are plotted as a function of photometric mass in
Fig. 17. These heights are good proxies to meteoroid structure,
although they depend to some extent on observational circum-
stances (e.g., range to the fireball) and on the slope of the tra-
jectory. Beginning heights show no dependence on mass and are
generally between 90 and 110 km. For consistency we use only
data from digital all-sky cameras in the plot. The two bright-
est fireballs were captured by the narrow-field cameras at higher
altitudes (see Table 1). On the other hand, both these fireballs
were located far from the all-sky cameras; the beginning of
EN 311015_180520 was 390 km from the closest camera and
the beginning of EN 311015_231301 was 270 km distant. If ob-
served from closer distances, the beginnings would lie somewhat
higher.
The maximum and end heights show large scatter. Many
fireballs exhibited multiple flares of similar brightness. Nev-
ertheless, there were differences in physical properties of the
meteoroids. This fact is mostly evident from the end heights.
There are differences of 25 km or more for meteoroids of sim-
ilar masses. The expected trend of deeper penetration for larger
bodies is only weakly present. The lowest end heights (below
50 km) were achieved by two quite small meteoroids. There are
no obvious differences in physical properties between different
branches of the stream.
Since the end height depends not only on the meteoroid prop-
erties but also on trajectory slope and entry speed, the PE crite-
rion (Ceplecha & McCrosky 1976), which compensates for these
effects, can be used to better evaluate meteoroid strengths. Ac-
cording to the PE criterion, meteoroids are classified into four
types: I, II, IIIA, and IIIB (Ceplecha 1988). Type I corresponds
to stony meteorites and type IIIB to soft cometary material. Fig-
ure 18 shows Taurid PE classification as a function of mass.
We can see that Taurids cover all four types, with a clear trend
of larger meteoroids being more fragile. Most of meteoroids
smaller than 30 grams belong to type II. Some meteoroids with
masses on the order of one gram clearly belong to type I. On
the other hand, most meteoroids above 30 gram belong to type
Article number, page 11 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fig. 13. Orientation of perihelia (latitude versus longitude) for Southern
Taurids observed in 2015. This plot was used to define the limits of the
new branch, as indicated by the ellipse. The fireballs that fell within
these limits but had different semimajor axes (outside the 7:2 resonance)
are plotted in purple. The fireballs outside these limits but within the
resonance are plotted in dark blue.
Fig. 15. Selected orbits of the Taurids from the new branch projected to
the plane of ecliptic.
Fig. 14. Semimajor axis as a function of solar longitude for Taurids
observed in 2015. For symbol explanation see Figs. 11 and 13. Error
bars are plotted for all fireballs. The extent of the 7:2 resonance with
Jupiter according to Asher & Clube (1993) is indicated.
Fig. 16. Number of fireballs as a function of photometric mass for
Northern Taurids, regular Southern Taurids, and the new branch.
IIIA or IIIB and only type IIIB is present above 300 gram. The
fact that the two largest meteoroids were very fragile was con-
firmed by fragmentation modeling (Sect. 4.1.1). Significant dif-
ferences between small and large meteoroids suggest the exis-
tence of some hierarchical structure and will be subject of future
studies.
Similar heterogeneity of Taurid physical properties was ob-
served recently by Matlovic et al. (2017). Brown et al. (2013)
reported a Taurid that penetrated down to 35 km and Madiedo
Article number, page 12 of 24
et al. (2014) and another one reaching 42.5 km. These authors
suggested that Taurids might drop meteorites. Our data do not
seem to support this possibility, since at least a ∼ 1 kg type I
Taurid meteoroid would be needed to produce any meteorites.
The SPMN 051010 fireball observed by Madiedo et al. (2014)
on October 5, 2010 had a semimajor axis 3.0 AU and perihelion
0.47 AU. It may not be Taurid at all. The SOMN 101031 fireball
observed by Brown et al. (2013), with a semimajor axis 2.9 AU,
was also not a typical Taurid.
All atmospheric and physical data are given in Table 5.
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Table 3. Orbital elements of asteroids discussed here as taken from the JPL database and converted to J2000.0 equinox.
Asteroid
2005 UR
2015 TX24
2005 TF50
2004 TG10
λ(cid:12)
216.44
218.81
219.60
223.32
a
2.254
2.269
2.272
2.234
e
0.882
0.872
0.869
0.862
q
0.266
0.290
0.298
0.308
i
6.94
6.05
10.70
4.18
ω
140.40
126.80
159.67
317.11
Ω
20.03
32.99
0.66
205.13
π
β
160.43
4.42
159.79
4.84
160.33
3.70
162.23 −2.84
Fig. 17. Fireball heights at beginning, end, and maximum light as a
function of photometric mass. Northern Taurids are plotted as dia-
monds, regular Southern Taurids as circles, and new branch members
as squares.
5. Related asteroids
We performed a search for asteroids with orbits similar to the
new Taurid branch responsible for the enhanced activity in 2015.
For that purpose, asteroids with q < 0.6 AU, 1.8 AU < a <
2.8 AU, and i < 12◦ were selected from the JPL Small-Body
Database2. There are 329 such asteroids known. We then plotted
selected orbital elements as a function of solar longitude at Earth
Minimum Orbit Intersection Distance (MOID) to be compared
with the observed fireballs. For fireballs we used solar longitude
at the time of impact as the independent variable. Since the aster-
oids did not impact Earth and their orbits do not intersect Earth's
orbit, we used for comparison the solar longitude, as seen from
the asteroid at the time when the asteroid is closest to the Earth's
orbit.
Figure 19 shows the comparison plot for eccentricity. We see
that there is nearly random distribution of asteroids with eccen-
tricities smaller than 0.84 in the solar longitudes of interest. At
higher eccentricities (0.86 – 0.88), however, there is a notice-
able concentration of four asteroids (2005 UR, 2015 TX24, 2005
TF50, and 2004 TG10) near solar longitude of 220◦. This con-
centration overlaps with the new Taurid branch. Moreover, it fol-
lows the same trend of decreasing eccentricity with increasing
solar longitude.
Other orbital elements are compared in Fig. 20. Perihelion
distance is basically a mirror image of eccentricity. Semimajor
axes of all four asteroids of interest fall within the Taurid branch
2 http://ssd.jpl.nasa.gov/sbdb_query.cgi, accessed January 25, 2017
Fig. 18. Value of PE criterion (Ceplecha & McCrosky 1976) as a func-
tion of photometric mass for all observed Taurids. Northern Taurids are
plotted as diamonds, regular Souther Taurids as circles, and new branch
members as squares. The dashed horizontal lines define the types I, II,
IIIA, and IIIB.
Fig. 19. Orbital eccentricity as a function for solar longitude at the clos-
est approach to the Earth's orbit for 2015 Taurid fireballs and asteroids
from JPL database. Asteroids, which are likely related to the new Tau-
rid branch are highlighted in magenta. Asteroids for which the relation
to the new branch was considered but not confirmed are shown as filled
rectangles. They may be related to other parts of the Taurid complex.
Article number, page 13 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fig. 20. Semimajor axis, inclination, perihelion distance, and longitude of perihelion as a function for solar longitude at the closest approach
to the Earth's orbit for 2015 Taurid fireballs and asteroids from JPL database. Asteroids, which are likely related to the new Taurid branch are
highlighted. The symbols are the same as in Fig. 19. The inclinations of Northern Taurids and asteroids with Ω > 180◦, which encounter the Earth
near their descending node in October/November, are plotted as negative.
range, i.e., also within the 7:2 resonance. As for inclination, only
2015 TX24 falls exactly within the Taurid branch range. The
2005 UR asteroid is somewhat off but only about a half degree
from the edge of the Taurid branch. However, Taurid fireballs
represent the part of the stream, which intersects Earth's orbit.
The whole stream is probably somewhat wider, so we consider
it likely that 2005 UR is also part of the stream. The 2005 TF50
asteroid matches all other elements very well but has an incli-
nation of 10.7◦, i.e., more than 4 degrees from the edge of the
Taurid branch. On the other hand, the orientation of perihelion is
not so far from the new Taurid branch (Fig. 22). But the orbit of
2004 TG10 is oriented in the opposite way relative to the eclip-
tic. This asteroid may be in fact related to Northern Taurids. At
least two asteroids, 2015 TX24 and 2005 UR, are therefore good
candidates for direct membership in the new branch of Southern
Taurids.
Asteroid 2015 TX24 was discovered by Pan-STARRS 1 on
October 8, 2015 and was observed for 18 days in October 2015.
It passed closest to the Earth's orbit on October 28, 2015, i.e.,
during the enhanced Taurid activity. The MOID of the Earth is
0.010 AU. The asteroid has an absolute magnitude of H = 21.5,
which corresponds to diameter 200 – 300 meters, assuming
albedo in the range 0.10 – 0.05.
Asteroid 2005 UR was discovered by the Catalina Sky Sur-
vey on October 23, 2005 and was observed for six days in Oc-
tober 2005. The MOID of the Earth is 0.034 AU. The absolute
magnitude is H = 21.6, i.e., very similar to that of 2015 TX24.
Asteroid 2005 UR approached the Earth's orbit at the end of De-
cember 2015 and was therefore only 17◦ in mean anomaly be-
hind the Taurids observed in 2015. Moreover, as noted by Olech
et al. (2016), Taurid activity was also enhanced when the asteroid
passed close to the Earth in October 2005.
Article number, page 14 of 24
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
theory of Asher & Clube (1993) and Asher & Izumi (1998) that
the meteoroids responsible for enhanced Taurid activity are in
7:2 resonance with Jupiter could be confirmed (at least for 2015
meteors). This fact cannot be revealed from lower precision data
such as those of Matlovic et al. (2017)3. Moreover, we found
that the Taurid branch, which is responsible for the enhanced ac-
tivity in 2015, forms an interesting orbital structure. Although
the enhanced activity lasted for 23 days according to our data,
all orbits had very similar orientation of the line of apsides, i.e.,
the longitude and latitude of perihelion. Since semimajor axes
were in a narrow range and all observed meteoroids had to in-
tersect Earth's orbit, only one free parameter remains. That is
why there is a good correlation between the longitude of the
ascending node (or, equivalently, solar longitude at the date of
observation), eccentricity, and perihelion distance.
There was, nevertheless, some spread of orbital elements
within the new branch. The longitudes of perihelia were within
the range 155.9 – 160◦ and latitudes of perihelia within the range
4.2 – 5.7◦. The semimajor axes were within the resonance lim-
its, 2.23 – 2.28 AU. The additional condition for the new branch
membership follows from the limited period of activity and can
be expressed, for example, in terms of eccentricity lying between
0.80 – 0.90. Three asteroids, 2015 TX24, 2005 UR, and 2005
TF50, fully or nearly satisfy all these conditions. Fig. 22 shows
three other asteroids (2003 WP21, 2007 UL12, and 2015 LM21)
with perihelia orientation not far from the new branch, but none
of them simultaneously fulfills both the semimajor axis and ec-
centricity criteria.
Since the Earth does not encounter the new branch every
year, it is evident that meteoroids of the new branch are not
spread along the whole orbit. The model and observations of
Asher & Clube (1993) and Asher & Izumi (1998) suggest that
the enhanced activity of Taurids is caused by a resonant swarm
of meteoroids, which extends ± 30 – 40◦ from the center of the
swarm in mean anomaly. It does not, however, necessarily mean
that the new branch observed in 2015 is identical to or represen-
tative of the whole swarm. The orbits of meteoroids observed by
the EN during the enhanced activity in 1995 had somewhat dif-
ferent characteristics than in 2015; these had larger semimajor
axes and smaller perihelia, which did not change so much with
solar longitude.
The new branch contains quite large bodies. Our brightest
fireball was caused by a body in excess of 1000 kg, which cor-
responds to diameter more than one meter, assuming that bulk
density was not higher than 2000 kg m−3. This body was disinte-
grated very high in the atmosphere and likely had high porosity
and low bulk density. The NASA JPL fireball page4 lists a fire-
ball with 10 times higher radiated energy, which occurred on
the same day (October 31, 2015 11:34:30 UT) above the Pa-
cific Ocean at a quite large height of 71 km. Considering the
unusual height, it is likely that that fireball belonged to the Tau-
rid new branch as well. The size of that body was 2-3 meters or
more. Two similar, slightly smaller, events occurred on Novem-
ber 2, 2005 (05:16:47 and 07:04:32), also over Pacific Ocean.
The heights of these bolides were 74 and 68.5 km, respectively.
These three fireballs are among the top five events with largest
heights among the 288 fireballs with known heights listed at the
NASA JPL page. Their trajectories and velocities are not given
but the Taurid radiant was above the horizon in all cases. We note
that 2005 was also a year of enhanced Taurid activity.
Fig. 21. Orbits of 2005 UR and 2015 TX24 in comparison with all Tau-
rids orbits from the new branch (gray).
Fig. 22. Comparison of perihelia orientation of 2015 Taurids with aster-
oids and comets from the JPL database.
The orbits of both 2005 UR and 2015 TX24 are plotted in
Fig. 21 together with the fireball orbits. There is a good overlap.
Orbital elements of all four asteroids discussed here are given in
Table 3. Asteroid 2004 TG10 is a large object with H = 19.4.
The albedo is very low and the diameter was estimated to be
1.40 ± 0.51 km (Nugent et al. 2015). The possible relation of
this asteroid to the Taurids was suggested already by Jenniskens
(2006), Porubcan et al. (2006), and Babadzhanov et al. (2008).
Asteroid 2005 TF50, with H = 20.3, is of intermediate size.
Its relation to comet 2P/Encke and the Taurids was proposed by
Porubcan et al. (2006) and Olech et al. (2016).
6. Discussion
We presented probably the most precise Taurid orbits obtained
to date. Thanks to the sufficient precision of semimajor axes, the
3 We observed 10 fireballs from their sample and their semimajor axes
are often off with respect to ours by several tenths of AU.
4 http://neo.jpl.nasa.gov/fireballs/, accessed February 3, 2017
Article number, page 15 of 24
A&A proofs: manuscript no. Spurny_Taurids
Fireball data therefore prove the presence of meter-sized
bodies among the Taurid new branch. Based on orbital similarity,
we argue that asteroids of several hundred meters in diameter are
members of the Taurid new branch as well. This is almost certain
for 2015 TX24, very likely for 2005 UR, and possible for 2005
TF50. We are not speaking about a distant relationship. The dis-
covered Taurid branch is simply a population of bodies with the
size range from several millimeters to several hundred of me-
ters, which all move together around the Sun. Every few years,
the Earth is encountering this branch for a period of about three
weeks. During that time, the chance of impact of an asteroid of
significant size (tens of meters) is significantly enhanced. Even
if intrinsically weak, bodies of such size can penetrate deep in
the atmosphere (Shuvalov & Artemieva 2002) and pose a hazard
to the ground.
We will allow theoretical celestial mechanicians to explain
the formation and evolution of the Taurid new branch and the
Taurid complex as a whole. A structure similar to the new branch
could be created by a disruption of a parent body at heliocentric
distance of about 3.6 AU (where the orbits come close together)
but ejection velocities up to 1.5 km s−1 and subsequent removal
of all non-resonant orbits would be needed. Also, asteroids 2005
TF50, 2015 TX24, and 2005 UR can all be related to 2004 TG10
but located at a different phase along the secular cycle as com-
puted for 2004 TG10 by Porubcan et al. (2006). Asteroid 2005
TF50 is about 2000 years behind, 2005 UR is about 2300 years
behind, and 2015 TX24 is about 2400 years behind 2004 TG10.
The elements ω, Ω, and i all agree well with this assumption. For
2005 TF50 and 2015 TX24 e and q are also in agreement. The
new Taurid branch can be also part of this relation. In fact, the or-
bital elements of the theoretical Southern Taurid meteors derived
from 2004 TG10, as computed by Babadzhanov et al. (2008), fall
perfectly among the Taurid branch fireballs in Fig. 12. Only in π
there is a difference of 2.5◦. But only the central part of the new
branch at λ(cid:12) ∼ 220◦ can be explained in this way.
7. Conclusions
We presented data of unprecedented precision for a large sample
of 144 Taurid fireballs observed by the European fireball net-
work in 2015. This data set contains precise and detailed data
on the Taurids covering 7 orders in mass, i.e., from tenths of a
gram to one-ton meteoroids. We have shown that the enhanced
Taurid activity in 2015 was produced by a well-defined branch
embedded within the much broader Southern Taurid stream. The
new branch can be characterized by the longitudes of perihelia
lying between 155.9 – 160◦, latitudes of perihelia between 4.2
– 5.7◦, semimajor axes between 2.23 – 2.28 AU, and eccentrici-
ties between 0.80 – 0.90. These orbits form a concentric ring in
the inner solar system with perihelia between 0.23 – 0.45 AU.
The new branch lies within the semimajor axis range spanned
by the 7:2 resonance, indicating strongly that the meteoroids re-
sponsible for the outbursts are within this resonance, as expected
from the model of Asher & Clube (1993). The Earth was the en-
countering members of the new branch at their ascending nodes
between October 25 and November 17. The orbital configuration
of the branch cause meteoroids with progressively lower eccen-
tricities, larger perihelion distances, and lower entry velocities to
encounter the Earth during the activity period.
The explanation of the structure and evolution of the new
branch and its relation to the whole Taurid complex must be left
to future theoretical studies. Nevertheless, we confirm earlier ob-
servations that the Taurid stream contains large meteoroids. This
is valid for the new branch in particular. The largest object we
Article number, page 16 of 24
observed was at least one meter in diameter. A ten times more
massive object observed on the same day over the Pacific Ocean
probably belonged to this new branch as well. Moreover, the or-
bits of asteroids 2015 TX24 and 2005 UR, both of diameters of
several hundreds of meters, place them within the new Taurid
branch as well. It is therefore very likely that the branch also
contains numerous objects of decameter size. Although our data
show that large Taurids have porous and fragile structure, objects
of tens or hundreds of meters in size pose a hazard to the ground
even if they have low intrinsic strength. Theoretical and obser-
vational studies and searches for related asteroids belonging to
this newly discovered and described branch of Southern Tau-
rids are therefore highly recommended. A better understanding
of this real source of potentially hazardous objects that are large
enough to cause significant regional or even continental damage
on the Earth is a task of capital importance.
Acknowledgements. This work was supported by the Praemium Academiae of
the Czech Academy of Sciences, grant 16-00761S from the Czech Science Foun-
dation, and the Czech institutional project RVO:67985815. Operation of the Slo-
vak station Stará Lesná was supported by the project ITMS No. 26220120029,
based on the supporting operational Research and development program financed
from the European Regional Development Fund. We thank especially J. Ke-
clíková, but also L. Shrbený and H. Zichová for a careful measuring of all pho-
tographic images. We also thank B. Pelc, T. Chlíbec, L. Sklenár and D. Šcerba
for their images of two brightest Taurids.
1436
References
Asher, D. J., Clube, S. V. M., & Steel, D. I., 1993, MNRAS 264, 93
Asher, D. J., & Clube, S. V. M., 1993, QJRAS 34, 481
Asher, D. J., & Izumi, K., 1998, MNRAS 297, 23
Babadzhanov, P. B., 2001, A&A 373, 329
Babadzhanov, P. B., Williams, I. P., & Kokhirova, G. I., 2008, MNRAS 386,
Borovicka, J., 1990, Bull. astr. Inst. Czechosl. 41, 391
Borovicka, J., Tóth, J., Igaz, A., et al., 2013, Meteorit. Planet. Sci. 48, 1757
Borovicka, J., Spurný, P., Grigore, V. I., & Svoren, J., 2017, Planet. Space Sci.,
in press, http://dx.doi.org/10.1016/j.pss.2017.02.006
Brown, P., Marchenko, V., Moser, D. E., Weryk, R., & Cooke, W., 2013, Mete-
orit. Planet. Sci. 48, 270
Ceplecha, Z., 1987, Bull. astr. Inst. Czechosl. 38, 222
Ceplecha, Z., 1988, Bull. astr. Inst. Czechosl. 39, 221
Ceplecha, Z. & McCrosky, R. E., 1976, J. Geophys. Res. 81, 6257
Ceplecha, Z., Spurný, P., Borovicka, J. & Keclíková, J., 1993, A&A 279, 615
Ceplecha, Z., Borovicka, J., Elford, W. G. et al., 1998, Space Sci. Rev. 84, 327
Clark, D. L., & Wiegert, P. A., 2011, Meteorit.Planet. Sci. 46, 1217
Clube, S. V. M., & Napier, W. M., 1984, MNRAS 211, 953
Dubietis, A., & Arlt, R., 2007, MNRAS 376, 890
Jenniskens, P., 2006, Meteor Showers and their Parent Comets. Cambridge Univ.
Press
Johannink, C., & Miskotte, K., 2006, WGN, J. of the IMO 34, 7
Madiedo, J. M., Ortiz, J. L., Trigo-Rodríguez, J. M. et al., 2014, Icarus 231, 356
Matlovic, P., Tóth, J., Rudawska, R., & Kornoš, L., 2017, Planet. Space Sci., in
press, http://dx.doi.org/10.1016/j.pss.2017.02.007
McBeath, A., 1999, WGN, J. of the IMO 27, 53
Napier, W. M., 2010, MNRAS 405, 1901
Nugent, C. R., Mainzer, A., Masiero, J., et al. 2015. ApJ 814, 117
Olech, A., ´Zoła¸dek, P., Wi´sniewski, M., et al., 2016, MNRAS 461, 674
Popescu, M., Birlan, M., Nedelcu, D. A., Vaubaillon, J., & Cristescu, C. P., 2014,
Porubcan, V., Kornoš, L., & Williams, I. P., 2006, Contrib. Astron. Obs. Skalnaté
A&A 572, A106
Pleso 36, 103
ReVelle, D. O., & Ceplecha, Z., 2001. In: Warmbein, B. (ed.), Proceedings of
the Meteoroids 2001 Conference, Kiruna, Sweden. ESA Special Publication
495, 507
Shiba, Y., 2016, WGN, J. of the IMO 44, 78
Shuvalov, V. V. & Artemieva, N. A., 2002, Planet. Space Sci. 50,181
Spurný, P., 1996, High fireball activity of the new subsystem of the Taurid com-
plex, ACM 1996, Versailles, poster presentation
Spurný, P., Borovicka, J., & Shrbený, L., 2007, in: G.B. Valsecchi, D. Vokrouh-
lický, & A. Milani (eds.) Near Earth Objects, our Celestial Neighbors: Op-
portunity and Risk, Proc. IAU Symposium 236 (Cambridge University Press),
p. 121
Tancredi, G., 2014, Icarus 233, 66
Whipple, F. L., 1940, Proceedings of the American Philosophical Society 83,
711
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Table 4. Radiant and orbital data for 2015 Taurid fireballs. The code of each fireball also contains the date (in ddmmyy format) and GMT time
corresponding to beginning rounded to whole second (in hhmmss format)
Code
Branch1
λ(cid:12)
EN231015_204348
EN231015_211327
EN241015_004546
EN241015_185031
EN251015_022301
EN251015_031725
EN261015_213736
EN261015_224031
EN271015_220749
EN281015_011855
EN301015_222401
EN311015_002325
EN311015_023900
EN311015_025717
EN311015_172431
EN311015_180520
EN311015_182902
EN311015_185530
EN311015_192126
EN311015_200534
EN311015_202117
EN311015_211904
EN311015_230919
EN311015_231301
EN011115_013625
EN011115_033911
EN011115_174410
EN011115_183646
EN011115_191104
EN011115_200918
EN011115_223909
S
N
N
S
SB
SB
SB
SB
SB
S
SB
N
SB
SB
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
N
SB
S
SB
SB
SB
209.908
209.928
210.075
210.825
211.138
211.176
212.933
212.977
213.951
214.084
216.958
217.040
217.134
217.147
217.749
217.777
217.794
217.812
217.830
217.861
217.872
217.912
217.988
217.991
218.091
218.176
218.763
218.799
218.823
218.864
218.968
αg
49.00
0.01
42.26
0.02
45.88
0.12
42.91
0.01
47.62
0.01
46.32
0.03
48.99
0.03
48.51
0.03
49.26
0.01
43.91
0.03
51.08
0.02
44.77
0.01
51.15
0.01
51.20
0.52
51.56
0.02
51.69
0.01
50.83
0.02
51.17
0.01
49.84
0.01
51.36
0.02
50.82
0.01
51.80
0.06
51.80
0.03
51.44
0.02
51.42
0.01
50.88
0.28
51.87
0.01
51.61
0.06
51.60
0.01
51.96
0.03
52.28
0.01
δg
15.31
0.02
20.04
0.02
18.83
0.08
11.61
0.01
13.76
0.04
13.59
0.02
14.44
0.03
13.96
0.03
13.94
0.05
11.12
0.07
14.52
0.06
19.16
0.01
14.34
0.03
14.38
0.04
14.50
0.03
14.59
0.02
14.50
0.03
14.68
0.03
13.99
0.01
14.57
0.07
14.66
0.01
14.74
0.02
14.69
0.02
14.49
0.04
14.58
0.03
20.51
0.06
14.84
0.02
15.34
0.06
14.15
0.01
14.48
0.06
14.64
0.02
vg
35.04
0.03
30.16
0.06
32.44
0.03
28.19
0.01
32.63
0.05
31.73
0.05
32.37
0.10
31.98
0.08
31.74
0.03
26.78
0.07
30.96
0.05
28.36
0.04
30.90
0.04
30.94
0.03
30.82
0.04
30.87
0.03
30.37
0.07
30.59
0.05
29.62
0.03
30.59
0.06
30.23
0.00
30.96
0.06
30.79
0.06
30.59
0.10
30.60
0.05
31.26
0.11
30.40
0.01
30.31
0.19
29.98
0.03
30.28
0.04
30.47
0.04
a
2.387
0.012
1.999
0.013
2.074
0.020
1.975
0.003
2.269
0.013
2.265
0.013
2.253
0.028
2.263
0.022
2.246
0.008
2.005
0.014
2.235
0.013
2.349
0.012
2.246
0.011
2.249
0.089
2.259
0.012
2.250
0.009
2.267
0.021
2.265
0.015
2.260
0.007
2.249
0.017
2.238
0.002
2.275
0.019
2.245
0.015
2.258
0.027
2.279
0.013
2.291
0.056
2.256
0.004
2.259
0.055
2.231
0.007
2.244
0.013
2.253
0.011
e
0.9249
0.0005
0.8578
0.0012
0.8913
0.0006
0.8272
0.0003
0.8955
0.0008
0.8847
0.0008
0.8921
0.0017
0.8874
0.0014
0.8838
0.0005
0.8057
0.0018
0.8733
0.0009
0.8436
0.0010
0.8726
0.0008
0.8732
0.0024
0.8719
0.0008
0.8724
0.0006
0.8665
0.0015
0.8693
0.0011
0.8563
0.0006
0.8689
0.0012
0.8640
0.0001
0.8743
0.0011
0.8713
0.0011
0.8689
0.0020
0.8696
0.0009
0.8793
0.0022
0.8665
0.0003
0.8657
0.0040
0.8599
0.0005
0.8643
0.0009
0.8671
0.0008
q
0.1792
0.0003
0.2842
0.0006
0.2255
0.0014
0.3414
0.0002
0.2370
0.0005
0.2612
0.0006
0.2430
0.0010
0.2549
0.0008
0.2610
0.0004
0.3896
0.0010
0.2832
0.0006
0.3674
0.0005
0.2861
0.0005
0.2850
0.0061
0.2893
0.0004
0.2872
0.0003
0.3026
0.0007
0.2960
0.0005
0.3247
0.0003
0.2949
0.0007
0.3044
0.0001
0.2860
0.0009
0.2889
0.0006
0.2960
0.0010
0.2971
0.0005
0.2765
0.0035
0.3013
0.0002
0.3034
0.0019
0.3127
0.0003
0.3044
0.0006
0.2994
0.0004
ω
134.92
0.03
303.17
0.03
309.90
0.24
116.78
0.02
127.73
0.04
124.80
0.06
127.03
0.07
125.55
0.06
124.85
0.04
111.24
0.08
122.22
0.05
292.16
0.02
121.83
0.04
121.95
1.03
121.40
0.04
121.69
0.02
119.81
0.03
120.59
0.03
117.29
0.02
120.78
0.06
119.71
0.02
121.73
0.13
121.49
0.06
120.62
0.06
120.41
0.03
302.85
0.56
120.00
0.02
119.74
0.13
118.77
0.02
119.68
0.07
120.23
0.03
i
5.39
0.04
5.02
0.04
2.46
0.14
5.51
0.01
6.26
0.06
5.54
0.03
5.66
0.05
5.99
0.05
6.19
0.07
5.73
0.07
5.65
0.08
2.41
0.01
5.89
0.04
5.86
0.24
5.76
0.05
5.71
0.02
5.31
0.03
5.28
0.04
5.34
0.02
5.51
0.09
5.06
0.01
5.56
0.04
5.58
0.03
5.62
0.06
5.49
0.03
2.69
0.11
5.22
0.02
4.46
0.08
5.86
0.01
5.68
0.08
5.64
0.03
π
164.84
0.03
153.09
0.03
159.94
0.24
147.62
0.02
158.88
0.04
155.99
0.06
159.97
0.07
158.54
0.06
158.82
0.04
145.34
0.08
159.19
0.05
149.17
0.02
158.98
0.04
159.11
1.03
159.17
0.04
159.48
0.02
157.62
0.03
158.42
0.03
155.13
0.02
158.65
0.06
157.60
0.02
159.65
0.13
159.50
0.06
158.62
0.06
158.51
0.03
160.99
0.56
158.77
0.02
158.55
0.13
157.61
0.02
158.56
0.07
159.21
0.03
Article number, page 17 of 24
A&A proofs: manuscript no. Spurny_Taurids
Table 4. continued.
Code
Branch1
λ(cid:12)
EN011115_234207
EN021115_020950
EN021115_021740
EN021115_022525
EN021115_024553
EN021115_182450
EN021115_195540
EN021115_201534
EN021115_205431
EN021115_213614
EN021115_215818
EN021115_220435
EN021115_232112
EN021115_234348
EN021115_235259
EN031115_002007
EN031115_011247
EN031115_012404
EN031115_025102
EN031115_031920
EN031115_193751
EN031115_195654
EN031115_202247
EN031115_204226
EN031115_212219
EN031115_212455
EN031115_213844
EN031115_221917
EN031115_221937
EN031115_222446
EN031115_225609
EN031115_230149
SB
SB
SB
SB
N
SB
SB
SB
N
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
N
SB
SB
SB
SB
SB
SB
SB
SB
SB
219.011
219.114
219.119
219.125
219.139
219.792
219.855
219.868
219.895
219.925
219.940
219.944
219.998
220.013
220.020
220.038
220.075
220.083
220.143
220.163
220.844
220.857
220.875
220.888
220.916
220.918
220.928
220.956
220.956
220.960
220.981
220.985
Article number, page 18 of 24
αg
52.47
0.03
51.90
0.02
51.96
0.02
52.64
0.03
47.24
0.05
51.84
0.39
52.70
0.02
52.22
0.03
52.44
0.02
52.39
0.02
52.49
0.02
52.45
0.01
52.14
0.01
50.43
0.04
52.65
0.01
52.51
0.02
53.06
0.02
52.76
0.02
52.88
0.02
52.25
0.02
52.95
0.07
53.26
0.04
51.46
0.07
52.73
0.03
52.99
0.04
53.50
0.02
52.78
0.11
52.88
0.02
52.99
0.01
52.00
0.04
53.44
0.02
53.14
0.03
δg
14.88
0.03
14.79
0.02
14.57
0.05
14.55
0.04
20.59
0.02
14.14
0.19
14.82
0.03
14.95
0.03
20.31
0.02
14.96
0.05
14.71
0.07
14.70
0.02
14.55
0.02
14.32
0.06
14.71
0.01
14.65
0.03
14.70
0.02
13.98
0.05
14.59
0.12
14.27
0.02
14.53
0.03
14.89
0.08
21.69
0.06
14.89
0.07
14.75
0.03
14.24
0.02
14.81
0.03
14.43
0.03
14.90
0.03
14.45
0.02
14.98
0.07
14.53
0.02
vg
30.64
0.16
30.26
0.04
30.19
0.05
30.60
0.05
27.81
0.04
29.61
0.04
30.26
0.02
29.97
0.09
30.88
0.05
30.03
0.07
30.05
0.05
29.99
0.05
29.78
0.04
28.59
0.11
30.02
0.03
29.83
0.03
30.22
0.06
30.02
0.09
30.06
0.04
29.67
0.12
29.73
0.03
29.91
0.10
29.92
0.10
29.61
0.04
29.75
0.09
29.89
0.08
29.60
0.05
29.50
0.09
29.71
0.03
29.05
0.08
30.02
0.05
29.72
0.09
a
2.263
0.043
2.278
0.012
2.262
0.014
2.259
0.013
2.121
0.010
2.262
0.064
2.274
0.005
2.272
0.027
2.248
0.013
2.270
0.021
2.272
0.014
2.265
0.013
2.277
0.012
2.247
0.029
2.250
0.008
2.232
0.009
2.249
0.017
2.278
0.023
2.253
0.013
2.269
0.030
2.276
0.014
2.261
0.029
2.262
0.030
2.273
0.011
2.278
0.027
2.254
0.021
2.272
0.023
2.251
0.025
2.268
0.009
2.273
0.022
2.277
0.015
2.269
0.026
e
0.8697
0.0032
0.8652
0.0009
0.8638
0.0011
0.8688
0.0009
0.8269
0.0009
0.8558
0.0022
0.8649
0.0003
0.8612
0.0020
0.8734
0.0009
0.8620
0.0015
0.8620
0.0010
0.8611
0.0010
0.8587
0.0009
0.8419
0.0026
0.8610
0.0006
0.8580
0.0007
0.8636
0.0013
0.8613
0.0018
0.8616
0.0009
0.8568
0.0025
0.8577
0.0007
0.8599
0.0022
0.8607
0.0021
0.8564
0.0008
0.8581
0.0020
0.8588
0.0016
0.8561
0.0013
0.8539
0.0020
0.8575
0.0007
0.8487
0.0017
0.8618
0.0011
0.8574
0.0020
q
0.2948
0.0016
0.3071
0.0005
0.3081
0.0007
0.2964
0.0006
0.3671
0.0007
0.3262
0.0046
0.3072
0.0003
0.3153
0.0010
0.2845
0.0005
0.3134
0.0008
0.3134
0.0006
0.3146
0.0005
0.3217
0.0005
0.3552
0.0013
0.3128
0.0003
0.3169
0.0004
0.3068
0.0007
0.3159
0.0010
0.3119
0.0008
0.3250
0.0014
0.3239
0.0008
0.3168
0.0011
0.3152
0.0012
0.3264
0.0006
0.3231
0.0010
0.3182
0.0008
0.3269
0.0014
0.3289
0.0010
0.3232
0.0004
0.3438
0.0009
0.3148
0.0007
0.3235
0.0010
ω
120.72
0.07
119.23
0.04
119.18
0.05
120.56
0.06
293.11
0.10
117.09
0.77
119.23
0.04
118.31
0.06
302.04
0.03
118.53
0.05
118.52
0.06
118.41
0.03
117.55
0.03
113.84
0.09
118.67
0.02
118.27
0.04
119.37
0.04
118.21
0.06
118.76
0.10
117.19
0.06
117.29
0.14
118.16
0.09
298.38
0.13
117.01
0.07
117.37
0.08
118.03
0.04
116.95
0.21
116.81
0.05
117.39
0.04
115.02
0.07
118.32
0.06
117.36
0.06
i
5.46
0.05
5.22
0.02
5.51
0.06
5.93
0.05
3.20
0.02
5.76
0.28
5.44
0.04
5.01
0.04
1.84
0.03
5.07
0.06
5.43
0.09
5.41
0.02
5.41
0.03
4.73
0.07
5.48
0.01
5.44
0.04
5.70
0.03
6.41
0.07
5.72
0.15
5.74
0.03
5.65
0.04
5.37
0.10
3.73
0.08
5.10
0.08
5.40
0.04
6.26
0.03
5.21
0.06
5.66
0.04
5.20
0.04
5.19
0.03
5.36
0.09
5.70
0.03
π
159.75
0.07
158.36
0.04
158.31
0.05
159.70
0.06
152.22
0.10
156.90
0.77
159.10
0.04
158.19
0.06
161.90
0.03
158.47
0.05
158.48
0.06
158.37
0.03
157.56
0.03
153.87
0.09
158.70
0.02
158.33
0.04
159.46
0.04
158.31
0.06
158.92
0.10
157.37
0.06
158.14
0.14
159.03
0.09
159.24
0.13
157.92
0.07
158.30
0.08
158.96
0.04
157.90
0.21
157.78
0.05
158.37
0.04
156.00
0.07
159.32
0.06
158.36
0.06
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Table 4. continued.
Code
Branch1
λ(cid:12)
EN031115_232829
EN031115_235911
EN041115_012728
EN041115_020201
EN041115_021111
EN041115_021452
EN041115_043317
EN041115_044559
EN041115_203853
EN041115_210403
EN041115_214032
EN041115_215226
EN041115_225243
EN041115_231355
EN051115_023102
EN051115_183559
EN051115_185259
EN051115_190203
EN051115_203651
EN051115_205304
EN051115_212802
EN051115_213128
EN051115_213433
EN051115_220108
EN051115_221253
EN051115_221501
EN051115_221906
EN051115_225625
EN051115_225852
EN051115_231201
EN051115_232719
EN051115_234939
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
S
SB
SB
SB
N
SB
SB
SB
SB
SB
N
SB
SB
S
SB
N
SB
221.004
221.025
221.087
221.111
221.117
221.120
221.216
221.225
221.888
221.905
221.931
221.939
221.981
221.996
222.133
222.804
222.816
222.823
222.889
222.900
222.924
222.927
222.929
222.947
222.956
222.957
222.960
222.986
222.988
222.997
223.007
223.023
αg
53.09
0.01
51.85
0.13
52.76
0.02
52.55
0.01
53.60
0.05
53.16
0.01
53.10
0.02
53.31
0.03
54.05
0.02
53.92
0.03
53.58
0.03
53.82
0.01
53.66
0.02
53.95
0.03
53.96
0.01
53.24
0.01
54.30
0.03
54.34
0.02
54.18
0.03
53.87
0.02
54.21
0.02
54.20
0.09
53.79
0.01
53.92
0.01
54.30
0.05
51.92
0.01
54.28
0.03
54.09
0.01
53.90
0.01
54.25
0.01
55.29
0.04
54.18
0.01
δg
15.30
0.02
16.41
0.18
14.08
0.01
14.48
0.02
14.77
0.03
14.64
0.01
14.11
0.02
14.43
0.02
14.80
0.05
15.16
0.11
14.86
0.02
14.86
0.01
14.61
0.03
14.39
0.05
14.96
0.03
13.90
0.02
14.56
0.03
14.66
0.02
14.44
0.03
21.71
0.01
14.94
0.04
14.90
0.16
14.60
0.02
14.47
0.01
14.77
0.02
21.71
0.01
15.20
0.02
14.54
0.04
15.19
0.01
15.06
0.00
21.85
0.03
14.47
0.01
vg
29.83
0.05
28.35
0.13
29.26
0.03
29.24
0.03
29.98
0.04
29.67
0.02
29.48
0.04
29.66
0.04
29.71
0.05
29.64
0.07
29.43
0.07
29.44
0.04
29.37
0.10
29.51
0.08
29.45
0.02
28.48
0.01
29.26
0.12
29.30
0.03
29.11
0.05
29.26
0.02
29.06
0.04
29.14
0.11
28.83
0.03
28.88
0.02
29.13
0.03
28.80
0.08
29.16
0.05
28.99
0.03
28.53
0.04
29.15
0.04
29.42
0.00
29.01
0.05
a
2.275
0.013
2.058
0.032
2.245
0.009
2.260
0.008
2.274
0.012
2.268
0.005
2.269
0.010
2.269
0.010
2.260
0.015
2.250
0.021
2.267
0.020
2.233
0.009
2.258
0.028
2.261
0.022
2.241
0.006
2.259
0.004
2.270
0.035
2.270
0.010
2.266
0.013
2.076
0.005
2.235
0.011
2.259
0.033
2.255
0.007
2.257
0.007
2.252
0.011
2.253
0.021
2.246
0.014
2.262
0.008
2.150
0.010
2.260
0.010
1.944
0.005
2.263
0.012
e
0.8595
0.0010
0.8324
0.0030
0.8503
0.0008
0.8508
0.0006
0.8610
0.0008
0.8568
0.0003
0.8539
0.0008
0.8564
0.0008
0.8569
0.0012
0.8559
0.0016
0.8535
0.0016
0.8526
0.0008
0.8523
0.0023
0.8541
0.0018
0.8530
0.0005
0.8398
0.0004
0.8510
0.0028
0.8516
0.0008
0.8487
0.0010
0.8461
0.0004
0.8475
0.0009
0.8492
0.0024
0.8448
0.0006
0.8455
0.0006
0.8488
0.0006
0.8454
0.0018
0.8493
0.0012
0.8471
0.0006
0.8373
0.0010
0.8495
0.0009
0.8445
0.0001
0.8473
0.0010
q
0.3196
0.0005
0.3449
0.0022
0.3360
0.0004
0.3372
0.0004
0.3160
0.0007
0.3250
0.0002
0.3315
0.0005
0.3258
0.0005
0.3234
0.0006
0.3242
0.0009
0.3321
0.0008
0.3292
0.0004
0.3335
0.0011
0.3300
0.0009
0.3295
0.0003
0.3618
0.0002
0.3382
0.0012
0.3371
0.0004
0.3427
0.0006
0.3195
0.0003
0.3409
0.0005
0.3405
0.0017
0.3499
0.0003
0.3487
0.0003
0.3404
0.0007
0.3484
0.0008
0.3384
0.0006
0.3458
0.0004
0.3498
0.0005
0.3400
0.0004
0.3023
0.0005
0.3454
0.0005
ω
117.78
0.02
115.81
0.27
116.01
0.04
115.83
0.03
118.20
0.09
117.19
0.02
116.44
0.04
117.10
0.06
117.40
0.06
117.34
0.10
116.37
0.06
116.83
0.03
116.24
0.05
116.64
0.07
116.77
0.03
113.02
0.02
115.65
0.05
115.78
0.03
115.15
0.06
298.63
0.04
115.48
0.05
115.43
0.21
114.38
0.02
114.50
0.02
115.47
0.11
294.60
0.02
115.72
0.06
114.81
0.04
114.83
0.02
115.48
0.02
301.24
0.08
114.86
0.02
i
4.77
0.02
2.78
0.21
5.95
0.02
5.40
0.02
5.65
0.05
5.55
0.02
6.10
0.03
5.85
0.02
5.64
0.06
5.13
0.13
5.30
0.03
5.40
0.02
5.61
0.04
6.02
0.06
5.32
0.03
5.93
0.02
5.79
0.03
5.70
0.02
5.83
0.03
2.98
0.01
5.26
0.04
5.31
0.19
5.43
0.02
5.64
0.02
5.50
0.03
3.32
0.02
5.00
0.03
5.65
0.05
4.74
0.01
5.15
0.01
2.86
0.05
5.76
0.01
π
158.80
0.02
156.87
0.27
157.12
0.04
156.95
0.03
159.33
0.09
158.33
0.02
157.67
0.04
158.34
0.06
159.30
0.06
159.26
0.10
158.32
0.06
158.79
0.03
158.24
0.05
158.64
0.07
158.91
0.03
155.84
0.02
158.49
0.05
158.62
0.03
158.06
0.06
161.50
0.04
158.42
0.05
158.38
0.21
157.32
0.02
157.47
0.02
158.44
0.11
157.53
0.02
158.70
0.06
157.81
0.04
157.84
0.02
158.49
0.02
164.22
0.08
157.90
0.02
Article number, page 19 of 24
A&A proofs: manuscript no. Spurny_Taurids
Table 4. continued.
Code
Branch1
λ(cid:12)
EN051115_235119
EN061115_001740
EN061115_002202
EN061115_003508
EN061115_005009
EN061115_011233
EN061115_011441
EN061115_011623
EN061115_025156
EN061115_030548
EN061115_040629
EN061115_164758
EN061115_174311
EN071115_015331
EN081115_010613
EN081115_033341
EN081115_181258
EN081115_202907
EN081115_212839
EN081115_234417
EN091115_001801
EN091115_003545
EN091115_011246
EN091115_011650
EN091115_032502
EN091115_041944
EN101115_212402
EN101115_235401
EN111115_004713
EN111115_031037
EN111115_181413
EN111115_181509
SB
SB
S
SB
S
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
N
SB
SB
S
SB
SB
SB
223.024
223.042
223.045
223.055
223.065
223.081
223.082
223.083
223.150
223.160
223.202
223.732
223.771
224.112
225.083
225.186
225.799
225.894
225.935
226.030
226.053
226.066
226.092
226.094
226.184
226.222
227.942
228.047
228.084
228.184
228.815
228.815
Article number, page 20 of 24
αg
54.53
0.01
54.44
0.01
52.54
0.02
54.24
0.03
54.31
0.04
54.27
0.01
53.57
0.09
54.04
0.02
54.42
0.22
54.69
0.02
54.36
0.01
54.85
0.00
54.88
0.01
54.80
0.04
55.57
0.04
55.71
0.04
55.87
0.01
55.47
0.02
55.15
0.01
55.75
0.03
55.15
0.02
55.43
0.05
55.46
0.04
55.36
0.02
55.99
0.02
56.51
0.03
56.75
0.03
57.00
0.04
56.33
0.01
57.37
0.12
57.27
0.06
57.11
0.03
δg
14.98
0.01
14.66
0.01
15.27
0.02
15.36
0.05
15.97
0.13
14.94
0.02
15.17
0.05
14.54
0.03
14.65
0.03
16.07
0.03
14.47
0.01
14.88
0.01
15.22
0.01
14.66
0.02
14.79
0.03
14.71
0.06
14.75
0.02
14.56
0.02
14.83
0.01
14.48
0.02
15.25
0.03
15.02
0.02
14.33
0.10
15.19
0.03
15.27
0.02
21.89
0.02
15.16
0.03
15.45
0.02
13.94
0.02
14.70
0.23
14.96
0.07
15.06
0.03
vg
29.35
0.08
29.23
0.03
28.14
0.09
29.15
0.07
29.26
0.12
29.08
0.04
28.72
0.10
28.91
0.12
29.17
0.08
29.54
0.05
29.01
0.01
28.94
0.01
29.14
0.04
28.71
0.12
28.62
0.12
28.57
0.11
28.31
0.01
28.02
0.07
27.80
0.02
28.02
0.08
27.82
0.16
27.91
0.03
27.72
0.11
27.87
0.12
28.27
0.02
29.90
0.06
27.47
0.05
27.66
0.03
27.01
0.08
27.64
0.06
27.30
0.08
27.20
0.00
a
2.274
0.023
2.275
0.007
2.261
0.025
2.256
0.018
2.251
0.033
2.256
0.010
2.263
0.029
2.266
0.032
2.279
0.041
2.276
0.014
2.263
0.004
2.240
0.002
2.279
0.012
2.260
0.030
2.271
0.032
2.257
0.028
2.264
0.003
2.274
0.021
2.260
0.007
2.254
0.021
2.272
0.041
2.261
0.011
2.237
0.028
2.258
0.029
2.277
0.007
2.387
0.017
2.242
0.015
2.260
0.009
2.256
0.020
2.248
0.026
2.266
0.025
2.262
0.004
e
0.8525
0.0018
0.8508
0.0006
0.8362
0.0022
0.8495
0.0014
0.8511
0.0027
0.8484
0.0008
0.8440
0.0022
0.8462
0.0027
0.8501
0.0021
0.8557
0.0011
0.8473
0.0003
0.8458
0.0002
0.8499
0.0009
0.8432
0.0027
0.8422
0.0028
0.8410
0.0025
0.8377
0.0003
0.8341
0.0018
0.8307
0.0006
0.8331
0.0019
0.8318
0.0038
0.8323
0.0008
0.8282
0.0026
0.8318
0.0028
0.8379
0.0006
0.8639
0.0012
0.8252
0.0014
0.8287
0.0007
0.8187
0.0020
0.8275
0.0017
0.8236
0.0021
0.8223
0.0002
q
0.3353
0.0009
0.3394
0.0003
0.3704
0.0011
0.3395
0.0008
0.3350
0.0015
0.3420
0.0004
0.3530
0.0015
0.3483
0.0014
0.3417
0.0027
0.3284
0.0007
0.3456
0.0002
0.3455
0.0001
0.3421
0.0004
0.3543
0.0014
0.3584
0.0014
0.3589
0.0014
0.3674
0.0002
0.3774
0.0008
0.3827
0.0003
0.3761
0.0009
0.3822
0.0017
0.3791
0.0007
0.3842
0.0014
0.3799
0.0013
0.3691
0.0004
0.3248
0.0007
0.3918
0.0007
0.3870
0.0006
0.4090
0.0009
0.3880
0.0019
0.3997
0.0011
0.4021
0.0003
ω
115.96
0.02
115.49
0.01
112.04
0.05
115.55
0.07
116.08
0.12
115.27
0.03
113.99
0.19
114.51
0.06
115.22
0.43
116.74
0.05
114.83
0.03
114.93
0.01
115.16
0.03
113.84
0.09
113.33
0.08
113.33
0.10
112.34
0.02
111.17
0.04
110.63
0.01
111.39
0.05
110.63
0.06
111.02
0.09
110.55
0.11
110.94
0.04
112.08
0.04
296.79
0.05
109.66
0.07
110.11
0.08
107.68
0.03
110.06
0.29
108.66
0.13
108.41
0.05
i
5.39
0.02
5.70
0.01
4.10
0.02
4.79
0.06
4.13
0.15
5.27
0.03
4.67
0.06
5.59
0.03
5.67
0.09
4.18
0.04
5.80
0.01
5.45
0.01
5.11
0.02
5.58
0.03
5.58
0.04
5.69
0.07
5.58
0.02
5.55
0.02
5.12
0.01
5.72
0.02
4.67
0.03
5.01
0.03
5.71
0.10
4.80
0.04
5.00
0.03
2.51
0.03
5.03
0.03
4.84
0.03
5.99
0.02
5.71
0.24
5.27
0.07
5.10
0.03
π
159.00
0.02
158.55
0.01
155.11
0.05
158.62
0.07
159.16
0.12
158.36
0.03
157.09
0.19
157.61
0.06
158.38
0.43
159.92
0.05
158.05
0.03
158.68
0.01
158.95
0.03
157.97
0.09
158.43
0.08
158.53
0.10
158.15
0.02
157.08
0.04
156.58
0.01
157.43
0.05
156.71
0.06
157.11
0.09
156.66
0.11
157.06
0.04
158.28
0.04
162.97
0.05
157.62
0.07
158.18
0.08
155.77
0.03
158.26
0.29
157.50
0.13
157.24
0.05
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Table 4. continued.
Code
Branch1
λ(cid:12)
EN111115_184540
EN111115_203917
EN111115_233243
EN121115_004717
EN121115_232341
EN131115_002058
EN131115_004858
EN131115_015008
EN131115_042559
EN161115_193458
EN161115_213048
EN161115_222246
EN171115_020907
EN171115_022102
EN231115_012005
EN231115_224311
EN281115_195251
N
SB
SB
SB
S
SB
SB
SB
SB
S
SB
S
SB
SB
N
S
S
228.837
228.916
229.037
229.089
230.037
230.077
230.097
230.139
230.248
233.906
233.987
234.023
234.182
234.190
240.202
241.102
246.038
αg
58.35
0.13
56.93
0.12
57.63
0.03
57.10
0.03
56.31
0.02
58.58
0.02
58.41
0.02
58.33
0.01
58.11
0.09
60.56
0.03
60.23
0.01
67.52
0.06
60.80
0.02
60.40
0.02
66.46
0.03
69.19
0.16
70.87
0.03
δg
22.60
0.02
14.97
0.04
15.13
0.01
15.23
0.04
14.51
0.05
15.58
0.06
15.35
0.02
14.98
0.01
15.37
0.02
14.63
0.01
15.09
0.01
16.82
0.02
15.58
0.08
15.51
0.02
24.65
0.01
16.03
0.02
16.55
0.02
vg
29.04
0.06
27.08
0.05
27.34
0.04
26.97
0.07
25.93
0.05
27.39
0.09
27.27
0.09
27.15
0.03
26.98
0.08
25.65
0.02
25.86
0.04
29.10
0.09
26.20
0.00
25.91
0.04
26.27
0.03
25.73
0.06
25.18
0.05
a
2.286
0.026
2.277
0.023
2.250
0.012
2.242
0.017
2.262
0.014
2.269
0.023
2.275
0.023
2.277
0.009
2.270
0.024
2.162
0.007
2.257
0.009
1.957
0.018
2.273
0.004
2.260
0.010
2.147
0.009
1.992
0.022
2.300
0.016
e
0.8494
0.0015
0.8212
0.0015
0.8235
0.0011
0.8183
0.0017
0.8041
0.0014
0.8251
0.0022
0.8236
0.0022
0.8219
0.0008
0.8195
0.0021
0.7942
0.0007
0.8025
0.0010
0.8372
0.0019
0.8083
0.0001
0.8036
0.0010
0.8035
0.0009
0.7853
0.0019
0.7936
0.0016
q
0.3443
0.0016
0.4072
0.0015
0.3970
0.0006
0.4074
0.0008
0.4431
0.0007
0.3968
0.0011
0.4013
0.0011
0.4057
0.0004
0.4097
0.0015
0.4450
0.0004
0.4457
0.0004
0.3185
0.0011
0.4357
0.0004
0.4440
0.0005
0.4219
0.0005
0.4277
0.0020
0.4747
0.0006
ω
294.88
0.25
107.77
0.24
109.02
0.06
107.90
0.07
103.82
0.05
108.96
0.06
108.42
0.05
107.93
0.03
107.51
0.17
104.03
0.06
103.50
0.02
119.17
0.12
104.54
0.06
103.68
0.03
286.63
0.06
106.76
0.31
99.91
0.05
i
2.78
0.03
5.09
0.05
5.19
0.02
4.85
0.04
5.00
0.04
4.94
0.06
5.09
0.02
5.41
0.01
4.91
0.03
5.66
0.01
5.20
0.01
6.26
0.03
4.94
0.08
4.84
0.02
2.93
0.01
5.83
0.04
5.12
0.02
π
163.68
0.25
156.71
0.24
158.08
0.06
157.01
0.07
153.88
0.05
159.05
0.06
158.54
0.05
158.09
0.03
157.77
0.17
157.95
0.06
157.51
0.02
173.21
0.12
158.74
0.06
157.89
0.03
166.80
0.06
167.88
0.31
165.97
0.05
1 N – Northern (13), S – Southern (18), SB – Southern new branch (113)
Article number, page 21 of 24
A&A proofs: manuscript no. Spurny_Taurids
Table 5. Physical data of 2015 Taurid fireballs. The entry velocity, heights of beginning, maximum brightness and end, average zenith distance of
the radiant, photometric mass, maximum absolute magnitude, PE coefficient, and classification according to PE are given. Code of each fireball
contains also date (in ddmmyy format) and GMT time corresponding to beginning rounded to whole second (in hhmmss format)
Code
Branch
EN231015_204348
EN231015_211327
EN241015_004546
EN241015_185031
EN251015_022301
EN251015_031725
EN261015_213736
EN261015_224031
EN271015_220749
EN281015_011855
EN301015_222401
EN311015_002325
EN311015_023900
EN311015_025717
EN311015_172431
EN311015_180520
EN311015_182902
EN311015_185530
EN311015_192126
EN311015_200534
EN311015_202117
EN311015_211904
EN311015_230919
EN311015_231301
EN011115_013625
EN011115_033911
EN011115_174410
EN011115_183646
EN011115_191104
EN011115_200918
EN011115_223909
EN011115_234207
EN021115_020950
EN021115_021740
EN021115_022525
EN021115_024553
EN021115_182450
EN021115_195540
EN021115_201534
EN021115_205431
EN021115_213614
EN021115_215818
EN021115_220435
EN021115_232112
EN021115_234348
EN021115_235259
EN031115_002007
EN031115_011247
EN031115_012404
EN031115_025102
EN031115_031920
EN031115_193751
EN031115_195654
EN031115_202247
EN031115_204226
EN031115_212219
EN031115_212455
S
N
N
S
SB
SB
SB
SB
SB
S
SB
N
SB
SB
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
N
SB
S
SB
SB
SB
SB
SB
SB
SB
N
SB
SB
SB
N
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
N
SB
SB
SB
v∞
km/s
36.98
32.30
34.25
30.58
34.30
33.40
34.38
33.95
33.75
28.87
32.98
30.38
32.63
32.63
33.05
33.06
32.61
32.80
31.90
32.76
32.43
33.08
32.78
32.56
32.41
32.92
32.64
32.55
32.24
32.47
32.52
32.60
32.05
32.00
32.38
29.72
31.93
32.46
32.19
33.00
32.17
32.16
32.11
31.80
30.66
31.98
31.78
32.10
31.87
31.85
31.43
31.99
32.13
32.12
31.84
31.92
32.04
Article number, page 22 of 24
Hbeg Hmax Hend
km
106.3
96.9
103.9
98.8
104.5
108.3
91.9
94.6
94.3
90.1
104.0
102.7
105.5
99.8
102.0
114.7
99.0
103.7
99.9
100.3
105.7
102.8
100.2
120.0
98.8
101.2
102.3
95.0
104.3
93.2
102.2
96.3
99.8
94.0
107.5
99.9
94.9
100.8
102.9
111.9
97.5
99.5
108.6
94.4
100.5
100.5
101.9
98.8
97.6
99.8
97.5
102.6
97.7
94.4
100.9
96.5
99.1
km
73.2
64.9
65.7
73.9
77.6
71.1
72.2
72.1
71.6
77.9
67.5
68.1
80.9
71.7
87.4
80.8
79.0
77.1
69.9
76.7
60.3
75.5
69.9
74.4
69.0
74.7
71.7
82.3
78.4
64.5
76.5
71.5
77.0
74.9
75.2
59.3
72.7
79.0
76.5
76.5
70.8
77.0
70.8
54.1
78.5
62.9
73.3
81.0
71.6
73.0
81.5
80.3
69.1
79.1
65.9
69.6
64.8
Zrad
km deg
59.7
54.7
38.2
57.8
31.4
64.8
67.3
65.6
69.7
44.6
52.3
66.9
43.4
64.0
38.2
61.3
41.9
62.4
65.8
42.8
38.9
65.5
32.0
59.3
47.5
68.5
53.3
57.9
81.9
76.5
72.1
57.6
69.8
74.0
63.7
69.4
69.5
62.6
52.2
67.2
53.5
55.2
46.8
71.3
33.6
64.1
57.3
36.9
41.0
58.2
54.0
71.6
75.6
67.2
80.1
65.8
63.4
73.8
53.2
61.3
37.3
62.8
31.3
57.6
69.7
45.1
45.1
68.2
45.3
63.3
47.5
53.9
62.7
71.2
55.8
75.1
53.6
72.1
42.9
73.8
45.4
60.7
73.3
40.8
39.3
63.6
35.5
48.4
32.8
70.9
58.9
33.2
35.1
58.7
37.9
67.4
41.9
61.1
48.4
69.1
71.3
56.1
57.8
69.9
54.6
68.4
44.5
72.8
51.0
59.2
68.1
42.1
41.7
59.8
Mass Mag
PE
Type
kg
0.0055
0.0021
0.0022
0.012
0.18
0.0011
0.0048
0.0041
0.20
0.0023
0.0047
0.0037
0.0055
0.21
0.0010
1300
0.0005
0.0006
0.0006
0.0020
0.0070
0.039
0.0006
34
0.021
0.0081
0.0089
0.0002
0.0049
0.0030
0.0021
0.0083
0.0020
0.0008
0.20
0.027
0.031
0.0018
0.0002
0.0003
0.15
0.0017
0.0010
0.015
0.0001
0.0031
0.0004
0.0003
0.0015
0.0004
0.0005
0.0053
0.0055
0.0008
0.0013
0.0017
0.013
-7.5
-4.2
-8.2
-6.3
-12.0
-4.9
-6.4
-6.5
-10.5
-5.0
-8.9
-6.3
-6.3
-11.5
-3.5
-18.6
-4.0
-3.0
-3.5
-5.1
-8.0
-10.2
-3.8
-15.8
-9.5
-9.4
-6.6
-2.4
-4.1
-6.9
-4.6
-7.5
-6.1
-4.2
-10.9
-7.8
-7.8
-5.3
-2.1
-3.2
-11.3
-6.7
-6.1
-9.3
-1.9
-7.5
-3.0
-2.6
-4.3
-3.0
-3.2
-6.4
-9.1
-4.2
-4.9
-6.7
-9.8
-4.52
-4.51
-4.89
-4.87
-5.87
-4.71
-4.90
-4.78
-5.52
-4.98
-5.04
-4.77
-5.17
-5.19
-5.12
-6.31
-4.66
-4.55
-4.62
-4.85
-4.43
-5.68
-4.61
-6.19
-4.92
-5.33
-4.62
-5.00
-5.00
-4.60
-4.72
-4.77
-5.09
-4.83
-5.57
-4.68
-4.76
-5.25
-4.73
-4.98
-5.37
-5.31
-4.68
-4.39
-4.78
-4.68
-4.28
-4.67
-4.59
-4.71
-4.80
-5.12
-5.09
-5.12
-4.37
-4.99
-4.92
I
I
II
II
IIIB
II
II
II
IIIA
II
II
II
II
II
II
IIIB
II
II
II
II
I
IIIB
II
IIIB
II
IIIA
II
II
II
II
II
II
II
II
IIIA
II
II
IIIA
II
II
IIIA
IIIA
II
I
II
II
I
II
II
II
II
II
II
II
I
II
II
P. Spurný et al.: Discovery of a new branch of the Taurid meteoroid stream
Mass Mag
PE
Type
Table 5. continued.
Code
Branch
EN031115_213844
EN031115_221917
EN031115_221937
EN031115_222446
EN031115_225609
EN031115_230149
EN031115_232829
EN031115_235911
EN041115_012728
EN041115_020201
EN041115_021111
EN041115_021452
EN041115_043317
EN041115_044559
EN041115_203853
EN041115_210403
EN041115_214032
EN041115_215226
EN041115_225243
EN041115_231355
EN051115_023102
EN051115_183559
EN051115_185259
EN051115_190203
EN051115_203651
EN051115_205304
EN051115_212802
EN051115_213128
EN051115_213433
EN051115_220108
EN051115_221253
EN051115_221501
EN051115_221906
EN051115_225625
EN051115_225852
EN051115_231201
EN051115_232719
EN051115_234939
EN051115_235119
EN061115_001740
EN061115_002202
EN061115_003508
EN061115_005009
EN061115_011233
EN061115_011441
EN061115_011623
EN061115_025156
EN061115_030548
EN061115_040629
EN061115_164758
EN061115_174311
EN071115_015331
EN081115_010613
EN081115_033341
EN081115_181258
EN081115_202907
EN081115_212839
EN081115_234417
EN091115_001801
SB
SB
SB
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
S
SB
SB
SB
N
SB
SB
SB
SB
SB
N
SB
SB
S
SB
N
SB
SB
SB
S
SB
S
SB
SB
SB
SB
S
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
SB
v∞
km/s
31.77
31.63
31.82
31.19
32.06
31.78
31.85
30.41
31.16
31.10
31.80
31.50
31.22
31.39
31.90
31.81
31.58
31.60
31.45
31.56
31.27
30.84
31.57
31.61
31.35
31.46
31.27
31.32
31.03
31.05
31.28
30.95
31.28
31.10
30.66
31.22
31.48
31.05
31.37
31.20
30.20
31.10
31.20
31.02
30.67
30.85
30.98
31.33
30.78
31.28
31.47
30.60
30.59
30.41
30.69
30.35
30.08
30.10
29.90
Hbeg Hmax Hend
km
101.5
101.6
98.9
100.8
95.9
99.1
98.6
98.1
106.5
98.1
100.2
102.8
101.8
100.1
95.0
98.7
94.1
93.9
97.3
92.3
98.0
99.9
99.7
99.6
101.1
103.5
96.5
91.4
105.9
99.3
102.4
107.8
96.7
99.9
98.7
105.3
101.0
102.5
99.3
104.8
102.2
100.4
94.2
98.9
98.9
97.2
94.5
98.8
101.1
104.0
99.6
104.4
92.4
96.4
99.1
100.2
103.1
99.4
102.9
km
72.7
83.8
64.6
72.9
72.0
74.5
69.5
76.1
63.7
63.9
71.7
68.4
83.0
82.7
74.9
64.7
68.2
64.2
67.9
71.7
65.7
68.3
79.6
78.5
71.7
70.0
72.3
70.8
62.2
70.8
64.7
77.0
76.3
76.7
73.3
72.0
71.0
61.5
73.7
68.5
62.9
66.6
74.4
60.1
83.6
69.0
77.5
75.9
71.1
79.7
79.8
76.7
78.7
72.3
65.9
68.1
68.3
70.9
77.6
Zrad
km deg
42.9
66.8
38.5
74.6
59.9
39.6
36.2
69.4
35.7
61.7
33.1
66.1
59.2
34.9
30.1
66.8
39.5
48.1
45.9
58.1
42.5
67.2
63.8
45.1
65.7
75.3
65.8
75.4
48.2
60.2
45.1
62.9
56.2
38.3
40.1
52.6
35.4
60.4
35.5
63.5
56.9
47.8
67.2
65.3
64.4
75.7
63.0
72.5
47.6
66.3
57.4
38.5
43.7
65.5
43.7
66.3
41.0
58.1
66.0
38.9
36.6
60.9
29.9
70.8
35.5
67.6
35.2
71.2
63.7
34.9
32.3
62.1
25.4
64.6
34.2
57.0
63.7
33.4
35.2
61.1
33.4
54.0
36.5
65.3
37.9
64.2
57.5
38.5
37.0
71.9
39.9
64.5
50.1
73.2
71.7
51.5
63.3
65.0
83.3
74.0
73.3
76.5
43.7
59.0
71.5
37.9
57.5
70.0
70.2
65.6
50.0
66.6
41.3
57.6
63.4
33.0
33.3
72.4
kg
0.0029
0.0003
0.0047
0.0003
0.0008
0.0010
0.0030
0.012
0.0018
0.0033
0.44
0.0076
0.0005
0.0005
0.21
0.031
0.016
0.0032
0.0015
0.019
0.011
0.0087
0.0004
0.0008
0.0010
0.093
0.0016
0.090
0.0038
0.050
0.010
0.0005
0.0006
0.0003
0.0019
0.11
0.0018
0.0026
0.0019
0.0044
0.0009
0.0052
0.0045
0.0031
0.0003
0.0004
0.0026
0.0020
0.0100
0.0098
0.0018
3.6
0.0014
0.0017
0.049
0.0020
0.0022
0.0062
0.0009
-6.5
-2.9
-7.7
-2.6
-3.8
-4.7
-7.0
-8.4
-4.1
-6.3
-12.7
-8.8
-3.0
-2.6
-10.6
-10.1
-9.8
-4.2
-5.6
-9.2
-8.8
-7.8
-3.3
-3.6
-6.3
-10.4
-4.7
-10.5
-7.4
-10.0
-9.6
-3.9
-4.0
-3.1
-5.7
-11.0
-6.2
-6.5
-6.8
-7.8
-4.3
-9.1
-7.8
-8.1
-3.4
-2.9
-5.0
-7.3
-8.2
-5.1
-4.2
-13.8
-5.9
-5.3
-9.8
-6.3
-5.0
-7.9
-5.2
-5.01
-5.08
-4.77
-4.78
-4.57
-4.85
-4.68
-5.40
-3.97
-4.56
-5.96
-5.01
-4.88
-4.89
-5.39
-5.21
-4.81
-4.29
-4.61
-5.25
-4.69
-4.80
-4.88
-4.85
-4.76
-5.20
-4.84
-5.62
-4.63
-5.54
-4.99
-5.02
-4.84
-4.90
-4.85
-5.52
-4.94
-4.56
-4.85
-4.86
-4.22
-5.11
-5.01
-4.58
-4.94
-4.57
-5.31
-5.15
-4.89
-4.64
-4.99
-5.92
-5.23
-4.95
-5.05
-4.90
-4.52
-5.08
-5.25
II
II
II
II
I
II
II
IIIA
I
I
IIIB
II
II
II
IIIA
II
II
I
II
IIIA
II
II
II
II
II
II
II
IIIA
II
IIIA
II
II
II
II
II
IIIA
II
I
II
II
I
II
II
I
II
I
IIIA
II
II
II
II
IIIB
II
II
II
II
I
II
II
Article number, page 23 of 24
Mass Mag
kg
0.041
0.0015
0.0003
0.0015
0.0052
0.011
0.12
0.022
0.91
0.23
0.27
0.0010
0.0010
0.014
0.0008
0.0005
0.0074
0.0003
0.0014
0.0009
0.0079
0.017
0.0004
0.0062
0.049
0.0040
0.074
0.0016
-9.3
-5.0
-3.3
-4.5
-8.5
-9.6
-10.9
-9.6
-13.5
-11.4
-11.8
-3.3
-3.8
-8.3
-5.4
-3.4
-8.2
-2.3
-5.0
-3.5
-6.1
-10.1
-3.6
-6.2
-10.6
-5.8
-10.3
-4.1
PE
Type
-4.85
-5.15
-4.96
-5.18
-4.95
-4.86
-5.78
-4.97
-5.92
-5.26
-5.70
-4.92
-5.29
-5.07
-5.17
-4.83
-5.25
-4.76
-4.63
-5.01
-4.58
-5.05
-4.86
-4.93
-5.16
-4.72
-5.25
-4.80
II
II
II
IIIA
II
II
IIIB
II
IIIB
IIIA
IIIB
II
IIIA
II
II
II
IIIA
II
II
II
I
II
II
II
II
II
II
II
A&A proofs: manuscript no. Spurny_Taurids
Hbeg Hmax Hend
km
96.2
88.7
98.4
99.6
99.3
102.4
98.1
97.3
95.0
97.2
98.5
102.0
93.5
98.9
96.9
95.7
97.4
94.5
103.1
93.7
101.5
100.4
97.9
93.5
95.1
93.2
96.8
108.0
Zrad
km deg
33.5
53.2
40.5
70.3
71.4
38.5
54.7
73.4
61.1
67.8
41.5
58.4
66.3
33.5
35.8
56.9
54.8
65.9
66.6
62.1
65.3
69.2
72.0
59.4
46.5
74.5
32.4
60.1
36.9
71.6
33.8
66.6
66.1
36.0
35.8
67.7
43.8
61.2
64.1
74.2
56.0
53.5
39.6
59.4
35.2
69.1
46.6
62.0
47.5
59.0
57.5
33.2
31.7
57.0
64.7
50.3
km
69.0
75.7
76.5
78.8
73.9
63.5
71.7
63.9
71.6
72.0
74.1
65.9
77.6
68.1
74.4
74.2
77.4
79.2
74.7
75.9
60.8
70.5
74.3
68.8
66.2
64.7
68.5
78.2
Table 5. continued.
Code
Branch
EN091115_003545
EN091115_011246
EN091115_011650
EN091115_032502
EN091115_041944
EN101115_212402
EN101115_235401
EN111115_004713
EN111115_031037
EN111115_181413
EN111115_181509
EN111115_184540
EN111115_203917
EN111115_233243
EN121115_004717
EN121115_232341
EN131115_002058
EN131115_004858
EN131115_015008
EN131115_042559
EN161115_193458
EN161115_213048
EN161115_222246
EN171115_020907
EN171115_022102
EN231115_012005
EN231115_224311
EN281115_195251
SB
SB
SB
SB
N
SB
SB
S
SB
SB
SB
N
SB
SB
SB
S
SB
SB
SB
SB
S
SB
S
SB
SB
N
S
S
v∞
km/s
29.97
29.73
29.88
30.13
31.62
29.77
29.75
29.11
29.55
29.76
29.66
31.36
29.47
29.51
29.07
28.20
29.48
29.34
29.17
28.90
28.19
28.28
31.25
28.26
27.98
28.37
28.06
27.74
Article number, page 24 of 24
|
1208.0377 | 1 | 1208 | 2012-08-01T23:33:55 | The anelastic equilibrium tide in exoplanetary systems | [
"astro-ph.EP"
] | Earth-like planets have anelastic mantles, whereas giant planets may have anelastic cores. As for the fluid parts of a body, the tidal dissipation of such solid regions, gravitationally perturbed by a companion body, highly depends on its internal friction, and thus on its internal structure. Therefore, modelling this kind of interaction presents a high interest to provide constraints on planet interiors, whose properties are still quite uncertain. Here, we examine the equilibrium tide in the solid central region of a planet, taking into account the presence of a fluid envelope. We first present the equations governing the problem, and show how to obtain the different Love numbers that describe its deformation. We discuss how the quality factor Q depends on the rheological parameters, and the size of the core. Taking plausible values for the anelastic parameters, and examinig the frequency-dependence of the solid dissipation, we show how this mechanism may compete with the dissipation in fluid layers, when applied to Jupiter- and Saturn-like planets. We also discuss the case of the icy giants Uranus and Neptune. | astro-ph.EP | astro-ph | SF2A 2012
S. Boissier, P. de Laverny, N. Nardetto, R. Samadi, D. Valls-Gabaud and H. Wozniak (eds)
2
1
0
2
g
u
A
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
7
7
3
0
.
8
0
2
1
:
v
i
X
r
a
THE ANELASTIC EQUILIBRIUM TIDE IN EXOPLANETARY SYSTEMS
F. Remus1,2,3, S. Mathis2, J.-P. Zahn1 and V. Lainey 3
Abstract.
Earth-like planets have anelastic mantles, whereas giant planets may have anelastic cores. As for the fluid parts of
a body, the tidal dissipation of such solid regions, gravitationally perturbed by a companion body, highly depends on its
internal friction, and thus on its internal structure. Therefore, modelling this kind of interaction presents a high interest to
provide constraints on planet interiors, whose properties are still quite uncertain.
Here, we examine the equilibrium tide in the solid central region of a planet, taking into account the presence of a
fluid envelope. We first present the equations governing the problem, and show how to obtain the different Love numbers
that describe its deformation. We discuss how the quality factor Q depends on the rheological parameters, and the size of
the core.
Taking plausible values for the anelastic parameters, and examinig the frequency-dependence of the solid dissipation,
we show how this mechanism may compete with the dissipation in fluid layers, when applied to Jupiter- and Saturn-like
planets. We also discuss the case of the icy giants Uranus and Neptune.
Keywords: planetary systems, dynamical evolution and stability
1 Introduction
Once a planetary system is formed, its dynamical evolution is governed by gravitational interactions between its com-
ponents, be it a star-planet or planet-satellite interaction. By converting kinetic energy into heat, the tides pertub their
orbital and rotational properties, and the rate at which the system evolves depends on the physical properties of tidal
dissipation. Therefore, to understand the past history and predict the fate of a binary system, one has to identify the
dissipative processes that achieve this conversion of energy. Planetary systems display a large diversity of planets, with
telluric planets having anelastic mantles and giant planets with possible anelastic cores (Udry & Santos 2007). Since the
tidal dissipation is closely related with the internal structure, one has to investigate its effects on each kind of materials
that may compose a planet. Studies have been carried out on tidal effects in fluid bodies such as stars and envelopes of
giant planets (Ogilvie & Lin 2004, 2007; Ogilvie 2009; Remus et al. 2012). However, the planetary solid regions, such
as the mantles of Earth-like planets or the rocky cores of giant planets may also contribute to tidal dissipation (see for
example Efroimsky 2012; Remus et al. 2012). We explore here the tidal dissipation in these solid parts of planets.
2 The system
Two-layer model. -- We will consider as a model a two-bodies system where the component A, rotating at the angular
velocity Ω, has a viscoelastic core of shear modulus µ, made of ice or rock, surrounded by a fluid envelope, such as an
ocean, streching out from core's surface (of mean radius Rc) up to planet's surface (of mean radius Rp). Both core and
envelope are considered homogeneous, with constant density ρc and ρo respectively. This model is represented on the left
panel of Fig. 1.
1 LUTH, Observatoire de Paris, CNRS, Université Paris Diderot, 92195 Meudon, France
2 Laboratoire AIM Paris-Saclay, CEA/DSM, CNRS, Université Paris Diderot, IRFU/SAp, 91191 Gif-sur-Yvette, France
3 IMCCE, Observatoire de Paris, CNRS, UPMC, USTL, 75014 Paris, France
© Société Francaise d'Astronomie et d'Astrophysique (SF2A) 2012
238
SF2A 2012
Configuration. -- We undertake to describe the tide exerted by B (of mass mB) on the solid core of A, when moving
in an elliptic orbit around A, with eccentricity e, at the mean motion ω. Since no assumption is made on the B's orbit,
we need to define an inclination angle I to determine the position of the orbital spin of B with respect to the total angular
momentum of the system (in the direction of ZR) wich defines an inertial reference plane (XR, YR), perpendicular to it.
The spin axis of A then presents an obliquity ε with respect to ZR. Refer to the right panel of Fig. 1 for a synthetic
representation of the system configuration.
Fig. 1. Left: the system is composed by a two-layer main component A, with an homogeneous and incompressible solid core and
an homogeneous static fluid envelope, and a point-mass perturber B orbiting around A. Right: B is supposed to move on an elliptical
orbit, inclined with respect to the inertial reference plane (XR, YR). The equatorial plane of A (XE , YE) is also inclined with respect to
this same reference plane.
To treat the complexity of the two-layer problem, we follow the methodology of Dermott (1979).
3 Tidal dissipation of the core in the case of a two-layer planet
Definition. -- The tidal perturbation exerted by B on the solid core of A results on one hand in its deformation, and
on the other hand in the dissipation of the tidal energy into heat leading to a lag angle δ between the line of centers and
the tidal bulge. This process can be modeled by the complex second-order Love number k2 defined as the ratio of the
perturbed gravific response potential Φ′ over the tidal potential U (Biot 1954, see also Tobie 2003 and Henning et al.
2009). Its real part represents then the purely elastic deformation of the potential of the core (Φ′) while its imaginary part
accounts for its anelastic tidal dissipation.
In practical calculations, we first have to develop U (and therefore Φ′) on spherical harmonics (Ym
2 ), each term having
a wide range of tidal frequencies σ2,m,p,q = (2 − 2p + q) ω − mΩ, for (m, p, q) ∈ [ − 2, 2] × [0, 2] × Z, resulting from the
expansion of U on the Keplerian elements using the Kaula transform (Kaula 1962, see also Mathis & Le Poncin-Lafitte
2009). Thus, the complex Love number k2 depends on the tidal frequency and the rheology of the core, and so does the
quality factor Q which quantifies the tidal dissipation (see for example Tobie 2003)
Q−1( ¯µ, σ2,m,p,q) = −
Im k2( ¯µ, σ2,m,p,q)
k2( ¯µ, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
19 µ
where the quantity ¯µ ≡ ¯µ1 + iii ¯µ2 =
the rheology) of the planet's core and its gravity gc.
2 ρc gc Rc
, where k2( ¯µ, σ2,m,p,q) =
Φ′( ¯µ, σ2,m,p,q)
U(σ2,m,p,q)
e−i[2δ(¯µ,σ2,m,p,q)] ,
(3.1)
k2(cid:12)(cid:12)(cid:12)
= (cid:12)(cid:12)(cid:12)
is the complex effective shear modulus, linked with the anelasticity (and thus
Case of a two-layer planet. -- Acting as an overload on the solid core, the fluid shell, deformed by the tide, modifies
both the tidal deformation and dissipation of the core. The second order Love number k2 takes then a different form than
in the fully-solid case
k2( ¯µ, σ2,m,p,q) =
1
(B + ¯µ1)2 + ¯µ2
2
×("(B + ¯µ1) C +
3
2α
¯µ1! +
3
2α
2# − iiiA D ¯µ2) ,
¯µ2
(3.2)
where α, A, B, C and D account for the planet's internal structure through the ratios of radii Rc
Rp
and densities ρo
ρc
.
The anelastic equilibrium tide in exoplanetary systems
Thus, the expression of the associated tidal dissipation rate
Q( ¯µ, σ2,m,p,q) = vuut1 +
9 ¯µ2(σ2,m,p,q)2
4α2 A2 D2
1 + (cid:16)B + ¯µ1(σ2,m,p,q)(cid:17) (cid:16) 2αC
3 + ¯µ1(σ2,m,p,q)(cid:17)
¯µ2(σ2,m,p,q)2
depends on the core's parameters (its size, density and rheological parameters) and the tidal frequency. Moreover, to
derive this expression of Q, no assumption has been made on the rheology of the core, except that it is linear under the
small tidal perturbations (i.e. core's material obeys the Hooke's law). Hence, it is valid for any linear rheological model.
Comparison with observations. -- To confront our model with observations, we need to introduce the global dissipation
factor, corresponding to a rescaling of the previous one to the planet surface and thus involving the second-order Love
number at the surface of the planet
239
(3.3)
2
Moreover, we need to choose a model to represent the way the core's material responds to the tidal perturbation, i.e.
a rheological model. Thus, from now on, we assume that the core behaves like a Maxwell body (see, for example, Tobie
2003).
Rc!5
Qeff = Rp
×(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
k2(Rp)
k2(Rc)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
× Q .
(3.4)
4 Application to giant planets
Application to gas giants. -- Using astrometric data covering more than a century, Lainey et al. (2009, 2012) suc-
ceeded in determining from observations the tidal dissipation in Jupiter and Saturn: namely, QJupiter = (3.56 ± 0.56) × 104
(Lainey et al. 2009), and QSaturn = (1.682 ± 0.540) × 103 determined by Lainey et al. (2012). Note that such high dissipa-
tion is required by the formation scenario of Saturn's system of Charnoz et al. (2011), in which the mid-sized satellites
are formed at the edge of the rings. These values, which seem to be in agreement with other observations related to
Jupiter's and Saturn's systems (see the corresponding references cited just above), are lower of up to one order of mag-
nitude than what was expected by previous formation scenarios (see, for example, Yoder & Peale 1981; Sinclair 1983),
and even lower than what the most up-to-date models of fluid tidal dissipation predict (see, for example, Ogilvie & Lin
2004; Wu 2005). Then, the question arises on the role of the possible solid central regions as sources of dissipation. Since
the composition of giant planets cores is poorly constrained (Guillot 2005), we explore in Fig. 2 the tidal dissipation of
Jupiter's and Saturn's core for a large range of values of the viscoelastic parameters considering the Maxwell rheological
model. The other parameters (planet and core sizes and masses) are indicated in the legend.
Fig. 2. Dissipation quality factor Qeff as a function of the viscoelastic parameters G and η, of a two-layer gas giant, using the Maxwell
model. Left: for a Jupiter-like planet at the tidal frequency of Io. Right: for a Saturn-like planet at the tidal frequency of Enceladus.
The red dashed line indicates the value of Qeff = {(3.56 ± 0.56) × 104, (1.682 ± 0.540) × 103} (for Jupiter and Saturn, respectively)
determined by Lainey et al. (2009, 2012). The blue lines corresponds to the lower and upper limits of the reference values taken by the
viscoelastic parameters G and η for an unknown mixture of ice and silicates. We assume the values of Rp = {10.97, 9.14} (in units of
R⊕), Mp = {317.8, 95.16} (in units of M⊕), Rc = {0.15, 0.26} × Rp, and Mc = {6.41, 18.65} (in units of M⊕).
240
SF2A 2012
In 2004, Ogilvie & Lin studied tidal dissipation in rotating giant planets resulting from the excitation by the tidal
potential of inertial waves in the convective region. Taking into account the presence of a solid core as a boundary
condition for the reflexion of inertial waves, they obtained a quality factor Qeff ≈ 5 × 105.
The present two-layer model proposes an alternative process that may reach the values observed in Lainey et al. (2009,
2012), depending on the viscosity η and the stiffness G.
To explain the tidal dissipation observed in the gas giant planets of our Solar System, all processes have to be taken
into account.
Application to ice giants. -- As in gas giants, the standard three-layer models for the interior structure of ice giants
predict the presence of a solid rocky core (see, for example, Hubbard et al. 1991; Podolak et al. 1995; Guillot 1999). But
it still remains an incertitude on the phase state of the intermediate "icy" layer located between the rocky core and the
convective atmosphere. Considering recent three-dimensional simulations of Neptune's and Uranus' dynamo that predict
that this region is a stably stratified conductive fluid one (Stanley & Bloxham 2004, 2006), Redmer et al. (2011) studied
the electric conductivity of warm dense water taking into account the phase diagram of water. Their results infer that part
of this shell is in the superionic state, i.e. a two-component system of both a conducting proton fluid and a crystalline
oxygen solid, and extends to about 0.42-0.56 of the planet radius. Thus, it seems reasonable to assume for our two-layer
model that the solid central region extends from the rocky core surface up to somewhere in the superionic shell.
We explore in Fig. 3 the tidal dissipation of Uranus' and Neptune's core for a large range of values of the viscoelastic
parameters, considering the Maxwell rheological model, for different core sizes.
Fig. 3. Dissipation quality factor Qeff as a function of the viscoelastic parameters G and η, of a two-layer ice giant, using the Maxwell
model. Top: for a Uranus-like planet at the tidal frequency of Miranda, with three different core sizes Rc = {0.12, 0.22, 0.32} × Rp. Bot-
tom: for a Neptune-like planet at the tidal frequency of Triton, with three different core sizes Rc = {0.14, 0.26, 0.32} × Rp. The red and
orange dashed lines indicate, respectively, the lowest and highest values of Qeff from formation scenarios:
Qeff = {5 × 103, 7.2 × 104}
for Uranus (Gavrilov & Zharkov 1977; Goldreich & Soter 1966) and Qeff = {9 × 103, 3.3 × 105} for Neptune (Zhang & Hamilton 2008;
Banfield & Murray 1992). The yellow dashed line indicates the value of Qeff from a study of Neptune's internal heat: Qeff = 1.7 × 102
(Trafton 1974). The blue lines corresponds to the lower and upper limits of the reference values taken by the viscoelastic parameters G
and η for an unknown mixture of ice and silicates. We assume the values of Rp = {3.98, 3.87} (in units of R⊕) and Mp = {14.24, 16.73}
(in units of M⊕). The core mass is obtained by integration of the density profiles of Helled et al. (2011) up to a given core size.
The anelastic equilibrium tide in exoplanetary systems
241
5 Dynamical evolution
Due to dissipation, the tidal torque has non-zero average over the orbit, and it induces an exchange of angular momentum
between each component and the orbital motion. This exchange governs the evolution of the semi-major axis a, the
eccentricity e of the orbit, the inclination I of the orbital plane, of the obliquity ε and that of the angular velocity of each
component (see for example Mathis & Le Poncin-Lafitte 2009). Depending on the initial conditions and on the planet/star
mass ratio, the system evolves either to a stable state of minimum energy (where all spins are aligned, the orbits are
circular and the rotation of each body is synchronized with the orbital motion) or the planet tends to spiral into the parent
star.
6 Conclusion
Our evaluations reveal a much higher dissipation in the solid cores of planets than that found by Ogilvie & Lin (2004)
for the fluid envelope of a planet having a small solid core. These results seem to be in good agreement with observed
properties of Jupiter's and Saturn's system (Lainey et al. 2009, 2012). In the case of the ice giants Uranus and Neptune,
too much uncertainties remain on internal structure to give an order of magnitude, other than a minimum value, of tidal
dissipation in the solid regions, which constitutes a first step in the tudy of such planets.
This work was supported in part by the Programme National de Planétologie (CNRS/INSU), the EMERGENCE-UPMC project EME0911, and the
CNRS Physique théorique et ses interfaces program.
References
Banfield, D. & Murray, N. 1992, Icarus, 99, 390
Biot, M. A. 1954, Journal of Applied Physics, 25, 1385
Charnoz, S., Crida, A., Castillo-Rogez, J. C., et al. 2011, Icarus, 216, 535
Dermott, S. F. 1979, Icarus, 37, 310
Efroimsky, M. 2012, ApJ, 746, 150
Gavrilov, S. V. & Zharkov, V. N. 1977, Icarus, 32, 443
Goldreich, P. & Soter, S. 1966, Icarus, 5, 375
Guillot, T. 1999, Science, 286, 72
Guillot, T. 2005, Annual Review of Earth and Planetary Sciences, 33, 493
Helled, R., Anderson, J. D., Podolak, M., & Schubert, G. 2011, ApJ, 726, 15
Henning, W. G., O'Connell, R. J., & Sasselov, D. D. 2009, ApJ, 707, 1000
Hubbard, W. B., Nellis, W. J., Mitchell, A. C., et al. 1991, Science, 253, 648
Kaula, W. M. 1962, AJ, 67, 300
Lainey, V., Arlot, J.-E., Karatekin, Ö., & van Hoolst, T. 2009, Nature, 459, 957
Lainey, V., Karatekin, Ö., Desmars, J., et al. 2012, ApJ, 752, 14
Mathis, S. & Le Poncin-Lafitte, C. 2009, A&A, 497, 889
Ogilvie, G. I. 2009, MNRAS, 396, 794
Ogilvie, G. I. & Lin, D. N. C. 2004, ApJ, 610, 477
Ogilvie, G. I. & Lin, D. N. C. 2007, ApJ, 661, 1180
Podolak, M., Weizman, A., & Marley, M. 1995, Planet. Space Sci., 43, 1517
Redmer, R., Mattsson, T. R., Nettelmann, N., & French, M. 2011, Icarus, 211, 798
Remus, F., Mathis, S., Zahn, J.-P., & Lainey, V. 2012, A&A, 541, A165
Sinclair, A. T. 1983, in Astrophysics and Space Science Library, Vol. 106, IAU Colloq. 74: Dynamical Trapping and Evolution in the
Solar System, ed. V. V. Markellos & Y. Kozai, 19 -- 25
Stanley, S. & Bloxham, J. 2004, Nature, 428, 151
Stanley, S. & Bloxham, J. 2006, Icarus, 184, 556
Tobie, G. 2003, PhD thesis, Université Paris 7 - Denis Diderot
Trafton, L. 1974, ApJ, 193, 477
Udry, S. & Santos, N. C. 2007, ARA&A, 45, 397
Wu, Y. 2005, ApJ, 635, 688
Yoder, C. F. & Peale, S. J. 1981, Icarus, 47, 1
Zhang, K. & Hamilton, D. P. 2008, Icarus, 193, 267
|
1207.4245 | 1 | 1207 | 2012-07-18T00:54:59 | Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system | [
"astro-ph.EP",
"astro-ph.IM"
] | We report the detection of UCF-1.01, a strong exoplanet candidate with a radius 0.66 +/- 0.04 times that of Earth (R_{\oplus}). This sub-Earth-sized planet transits the nearby M-dwarf star GJ 436 with a period of 1.365862 +/- 8x10^{-6} days. We also report evidence of a 0.65 +/- 0.06 R_{\oplus} exoplanet candidate (labeled UCF-1.02) orbiting the same star with an undetermined period. Using the Spitzer Space Telescope, we measure the dimming of light as the planets pass in front of their parent star to assess their sizes and orbital parameters. If confirmed, UCF-1.01 and UCF-1.02 would be called GJ 436c and GJ 436d, respectively, and would be part of the first multiple-transiting-planet system outside of the Kepler field. Assuming Earth-like densities of 5.515 g/cm^3, we predict both candidates to have similar masses (~0.28 Earth-masses, M_{\oplus}, 2.6 Mars-masses) and surface gravities of ~0.65 g (where g is the gravity on Earth). UCF-1.01's equilibrium temperature (T_{eq}, where emitted and absorbed radiation balance for an equivalent blackbody) is 860 K, making the planet unlikely to harbor life as on Earth. Its weak gravitational field and close proximity to its host star imply that UCF-1.01 is unlikely to have retained its original atmosphere; however, a transient atmosphere is possible if recent impacts or tidal heating were to supply volatiles to the surface. We also present additional observations of GJ 436b during secondary eclipse. The 3.6-micron light curve shows indications of stellar activity, making a reliable secondary eclipse measurement impossible. A second non-detection at 4.5 microns supports our previous work in which we find a methane-deficient and carbon monoxide-rich dayside atmosphere. | astro-ph.EP | astro-ph | ACCEPTED BY APJ 2012-06-07
Preprint typeset using LATEX style emulateapj v. 8/13/10
TWO NEARBY SUB-EARTH-SIZED EXOPLANET CANDIDATES IN THE GJ 436 SYSTEM
KEVIN B. STEVENSON, JOSEPH HARRINGTON, AND NATE B. LUST
Planetary Sciences Group, Department of Physics, University of Central Florida
Orlando, FL 32816-2385, USA
Department of Planetary Sciences and Lunar and Planetary Laboratory, The University of Arizona, Tucson, AZ 85721, USA
NIKOLE K. LEWIS
European Organisation for Astronomical Research in the Southern Hemisphere (ESO), Casilla 19001, Santiago 19, Chile
GUILLAUME MONTAGNIER
Space Science Institute, 4750 Walnut St, Suite 205, Boulder, CO 80301, USA
JULIANNE I. MOSES
Southwest Research Institute, 1050 Walnut St., Suite 300, Boulder CO, 80302, USA
CHANNON VISSCHER
JASMINA BLECIC, RYAN A. HARDY, PATRICIO CUBILLOS, AND CHRISTOPHER J. CAMPO
Planetary Sciences Group, Department of Physics, University of Central Florida
Orlando, FL 32816-2385, USA
Accepted by ApJ 2012-06-07
ABSTRACT
We report the detection of UCF-1.01, a strong exoplanet candidate with a radius 0.66 ± 0.04 times that of
Earth (R⊕). This sub-Earth-sized planet transits the nearby M-dwarf star GJ 436 with a period of 1.365862 ±
8×10- 6 days. We also report evidence of a 0.65 ± 0.06 R⊕ exoplanet candidate (labeled UCF-1.02) orbiting
the same star with an undetermined period. Using the Spitzer Space Telescope, we measure the dimming of
light as the planets pass in front of their parent star to assess their sizes and orbital parameters. If confirmed,
UCF-1.01 and UCF-1.02 would be called GJ 436c and GJ 436d, respectively, and would be part of the first
multiple-transiting-planet system outside of the Kepler field. Assuming Earth-like densities of 5.515 g/cm3,
we predict both candidates to have similar masses (∼0.28 Earth-masses, M⊕, 2.6 Mars-masses) and surface
gravities of ∼0.65 g (where g is the gravity on Earth). UCF-1.01's equilibrium temperature (T eq, where emitted
and absorbed radiation balance for an equivalent blackbody) is 860 K, making the planet unlikely to harbor life
as on Earth. Its weak gravitational field and close proximity to its host star imply that UCF-1.01 is unlikely
to have retained its original atmosphere; however, a transient atmosphere is possible if recent impacts or tidal
heating were to supply volatiles to the surface. We also present additional observations of GJ 436b during
secondary eclipse. The 3.6-µm light curve shows indications of stellar activity, making a reliable secondary
eclipse measurement impossible. A second non-detection at 4.5 µm supports our previous work in which we
find a methane-deficient and carbon monoxide-rich dayside atmosphere.
Subject headings: planetary systems - stars: individual: GJ 436 - techniques: photometric
2
1
0
2
l
u
J
8
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
4
2
4
.
7
0
2
1
:
v
i
X
r
a
1. INTRODUCTION
The search for Earth-sized planets around main-sequence
stars has progressed expeditiously in the last year. Recent
discoveries include two Earth-sized planets (0.868 and 1.03
Earth radii, R⊕) from the Kepler-20 system (Fressin et al.
2012), two planet candidates (0.759 and 0.867 R⊕) from the
KIC 05807616 system (Charpinet et al. 2011), and a three-
planet system (0.78, 0.73, and 0.57 R⊕) orbiting KOI-961
(Muirhead et al. 2012).
The search for a second planet in the GJ 436 system be-
gan shortly after the transit detection and confirmed eccentric
orbit of GJ 436b (Gillon et al. 2007; Deming et al. 2007). In
2008, a ∼5-M⊕ planet on a 5.2-day orbit was proposed by
[email protected]
Ribas et al. (2008, later retracted) due to three lines of evi-
dence. First, the lack of detectable GJ 436b transits at the
time of its 2004 discovery using radial-velocity (RV) mea-
surements (Butler et al. 2004) suggests a change in orbital
inclination due to a perturber. Second, given a circulariza-
tion timescale of ∼30 Myr (Deming et al. 2007) and the es-
timated 6-Gyr age of the system (Torres 2007), GJ 436b's
non-circular orbit suggests another planet is pumping up its
eccentricity. Third, there was evidence of a residual low-
amplitude RV signal in a 2:1 mean-motion resonance with
GJ 436b (Ribas et al. 2008). The inferred planet was discred-
ited by orbital-dynamic simulations (Bean & Seifahrt 2008;
Demory et al. 2009) and the absence of transit timing varia-
tions (TTVs) with two transit events with the Near Infrared
Camera and Multi Object Spectrograph camera on the Hub-
2
Stevenson et al.
ble Space Telescope (Pont et al. 2009) and over a 254-day
span using ground-based H-band observations (Alonso et al.
2008).
Ballard et al. (2010b)'s analysis of 22 days of nearly contin-
uous observations of GJ 436 during NASA's EPOXI mission
ruled out transiting exoplanets >2.0 R⊕ outside GJ 436b's
2.64-day orbit (out to a period of 8.5 days) and >1.5 R⊕ in-
terior to GJ 436b, both with a confidence of 95%. Aided
by a ∼70-hour Spitzer observation of GJ 436 at 8.0 µm,
Ballard et al. (2010a) postulated the presence of a 0.75-R⊕
planet with a period of 2.1076 days. However, the predicted
transit was not detected in an 18-hour follow-up observa-
tion with Spitzer at 4.5 µm. The candidate transit signals
in the EPOXI data were likely the result of correlated noise
(Ballard et al. 2010a).
In this paper we present Spitzer primary-transit observa-
tions of UCF-1.01 and UCF-1.02 at 4.5 µm (including an
independent analysis), a phase curve of GJ 436b at 8.0 µm
in which transits of UCF-1.01 are modeled, and a publicly-
available EPOXI light curve phased to the period of UCF-
1.01. We also include secondary-eclipse observations of GJ
436b at 3.6 and 4.5 µm.
In Section 2, we describe the observations and data analy-
sis. Section 3 presents Time-series Image Denoising (TIDe, a
wavelet-based technique used to improve image centers) and
provides an example analysis using a fake dataset. In Sec-
tion 4, we discuss the specific steps taken with each of the six
Spitzer datasets, the details of our independent analysis, and
transit results from the EPOXI light curve. Section 5 describes
how we eliminate false positives, our radial-velocity analysis,
mass constraints on both sub-Earth-sized exoplanets, and or-
bital and atmospheric constraints on UCF-1.01. Finally, we
give our conclusions in Section 6 and supply the full set of
best-fit parameters with uncertainties in the Appendix.
2. OBSERVATIONS AND DATA ANALYSIS
2.1. Observations
We observed GJ 436 at 3.6 and 4.5 µm using Spitzer's In-
fraRed Array Camera (Fazio et al. 2004). Including the two
previously analyzed data sets listed in Table 1, we present six
Spitzer observations spanning just over three years.
2.2. POET Pipeline and Modeling
Our Photometry for Orbits, Eclipses, and Transits (POET)
pipeline produces systematics-corrected light curves from
Spitzer Basic Calibrated Data. We flag bad pixels, calcu-
late image centers from a Gaussian fit, and apply interpolated
aperture photometry (Harrington et al. 2007) with a broad
range of aperture sizes in 0.25-pixel increments. To achieve
more precise image centers in the 2010 January 28 dataset, we
utilize TIDe (see Section 3). For a more detailed description
of POET, see Campo et al. (2011) and Stevenson et al. (2011).
We model the light curve as follows:
F(x,y,t) = FsE(t)R(t)S(t)M(x,y),
(1)
where F(x,y,t) is the measured flux centered at position (x,y)
on the detector at time t; Fs is the (constant) system flux out-
side of transit events; E(t) is the primary-transit or secondary-
eclipse model component; R(t) = 1 - e- r0t+r1 + r2(t - r3) is the
time-dependent ramp model component with free parameters
r0 - r3; S(t) = s0 cos[2π(t - s1)/p] is the phase variation at 8.0
The
small-planet
uniform-source
µm with free parameters s0 and s1 and p being the fixed pe-
riod of GJ 436b; and M(x,y) is the Bilinearly-Interpolated
Subpixel Sensitivity (BLISS) map. We follow the method de-
scribed by Stevenson et al. (2011) when determining the opti-
mal bin sizes of the BLISS maps.
and
equations
(Mandel & Agol 2002) describe the secondary-eclipse and
primary-transit model components. We apply a non-linear
stellar limb-darkening model (Claret 2000; Beaulieu et al.
2008)
to UCF-1.01 transits with coefficients a1-a4 =
(0.79660, -1.0250, 0.82228, -0.26800). Spitzer data has well
documented systematic effects that our Levenberg-Marquardt
minimizer fits simultaneously with the transit/eclipse pa-
rameters. BLISS mapping (Stevenson et al. 2011) models
the position-dependent systematics (such as intrapixel vari-
ability and pixelation) and linear or asymptotically constant
exponential functions model the time-dependent systematics.
A Metropolis Random-Walk Markov-chain Monte Carlo
(MCMC) algorithm assesses the uncertainties (Campo et al.
2011). Each MCMC run begins with a least-squares mini-
mization, a rescaling of the Spitzer-supplied uncertainties so
that the reduced χ2 = 1, and a second least-squares minimiza-
tion using the new uncertainties. We test for convergence ev-
ery 105 steps, terminating only when the Gelman & Rubin
(1992) diagnostic for all free parameters has dropped to
within 1% of unity using all four quarters of the chain. We
also examine trace and autocorrelation plots of each parame-
ter to confirm convergence visually. We estimate the effective
sample size (ESS, Kass et al. 1998) and autocorrelation time
for each free parameter and apply the longest autocorrelation
time from each event to determine the number of steps be-
tween independent samples in each MCMC chain. We place
a prior on UCF-1.01's semi-major axis (a/R⋆ = 9.1027+0.0060
- 0.0067)
by applying its known period and GJ 436b's well constrained
semi-major axis and period (Knutson et al. 2011) to Kepler's
third law. Without a prior, the uncertainties for the semi-major
axis (and any correlated parameters) are larger, but not unsta-
ble. We also place a flat prior on UCF-1.02's ingress/egress
time of < 0.1 hours because it is unconstrained by the data.
3. TIME-SERIES IMAGE DENOISING
Here we describe an application of wavelets that improves
image centering, resulting in more precise aperture photome-
try and better handling of the position-dependent systematics.
Readers primarily interested in the science results can skip to
Section 4.
3.1. Introduction
Photon noise in short exposures can cause significant shifts
between the fitted and real stellar centers. With imprecise
centering over multiple frames, varying amounts of light fall
within the improperly placed apertures, thus increasing light-
curve scatter. The sensitivity to precise centering increases
with smaller aperture sizes, causing a given change in aper-
ture position to produce larger changes in flux. To improve
centering, one could sum many exposures, but wavelet filter-
ing allows the same noise reduction over a shorter time span.
This is important because Stevenson et al. (2010) detect 0.04-
pixel (0.05 arcsec) pointing variations for IRAC data over ∼5
seconds at > 10σ, which limits the span of an averaging win-
dow to a few seconds. Our wavelet filter is called Time-series
Image Denoising (TIDe, pronounced "tidy"). It affects only
high-frequency components, such as photon noise, without af-
Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system
3
TABLE 1
OBSERVATION INFORMATION
Total Frames
Wavelength
Observation Date Duration
[minutes]
Frame Time
[seconds]
2008 July 14
2010 January 28
2010 June 29
2011 January 24
2011 February 1
2011 July 30
[µm]
8.0
4.5
4.5
4.5
3.6
4.5
aK10 = parts were published by Knutson et al. (2010), B10 = Ballard et al. (2010a).
588,480
488,960
49,536
51,712
168,576
36,160
4,207
1,081
363
369
369
258
0.4
0.1
0.4
0.4
0.1
0.4
Spitzer
Pipeline
S18.18.0
S18.18.0
S18.18.0
S18.18.0
S18.18.0
S18.18.0
Previous
Publications1
K10
B10
–
–
–
–
fecting low-frequency components like transits or eclipses. It
retains the time resolution of the data.
In addition to improving aperture photometry, precise cen-
tering (see example in Section 3.3) improves our ability to
model and remove position-dependent systematics accurately,
for example by reducing the smallest meaningful bin size for
BLISS mapping. TIDe does correlate the data in time, which
complicates error analysis and makes it computationally in-
tense because the correlation depends on the signal and thus
varies in time. So, we use TIDe-cleaned images only for cen-
tering (whose uncertainties do not propagate), and perform
photometry on the unfiltered images.
As with a windowed (sliding) Fourier transform (WFT),
wavelets decompose a signal into independent contributions
at each scale and location (similar to frequency and time)
within the signal. As an example, the Fourier transform of
a piece of music can discern the average pitch and timbre of
all the instruments, but wavelets can identify individual notes
and the instruments that played them at any given time. The
wavelet transform of a univariate time series thus has two di-
mensions, for location and scale. Unlike the WFT, wavelets
do not suffer from a fixed resolution (or window size), so
they retain both good temporal resolution for high-frequency
events and good scale resolution for low-frequency events.
Torrence & Compo (1998) provide an accessible introduction
to wavelets.
TIDe's improvement in precision and benefits to the light-
curve fit can vary based on the source brightness, aperture
size, BLISS map resolution, etc. This method is applicable
to most photon-noise-limited photometric observations where
the cadence is significantly shorter than the duration or period
of the time-varying object of interest.
3.2. Description of TIDe
TIDe applies discrete wavelet denoising independently to
multiple time series, each comprised of the values measured
in a single pixel as a function of time (i.e., frame number).
Every pixel is associated with such a time series, and each
one is denoised independently of adjacent image pixels. The
transformed data (known as wavelet coefficients) for each
pixel time series have a location (or time) dimension and
a scale (or level) dimension. The wavelet coefficients map
the discrete wavelet to the data at each scale and instant in
time. The lowest level (or finest scale) of decomposition in
a wavelet transform describes how the data change on the
shortest timescales. Assuming that this level is dominated by
noise, we can eliminate wavelet coefficients with magnitudes
below a certain threshold (hard thresholding) or merely atten-
uate them (soft thresholding) to reduce their contribution to
the overall signal. Adjusting a collection of estimates together
in this way can be shown to improve the average quality of the
estimates by introducing a small bias that is more than com-
pensated for by reduced variance. These techniques can also
be applied to successively higher levels, but they have less
impact at longer timescales where the signal dominates over
the noise. After thresholding, we recombine all of the ad-
justed wavelets to generate a less-noisy version of the original
pixel time series. For each frame, an image is re-created from
the many denoised time series, and centering is performed us-
ing that image. There is no explicit spatial denoising, but to
the extent that there are spatial correlations between images
at different epochs, there is an implicit spatial denoising in
the processed image that improves center estimation. The ef-
fectiveness of TIDe is determined by the threshold at which
wavelet coefficients are zeroed, the type of thresholding tech-
nique applied, and the number of levels to which the method
is applied (Donoho & Johnstone 1994; Chang et al. 2000).
Various wavelet thresholding techniques exist, each with its
own advantages and disadvantages. Two common methods
for suppressing noise are hard and soft universal thresholding
and are defined, respectively, as follows:
ω = yI(y > T ),
ω = sgn(y)(y - T)I(y > T ),
(2)
(3)
where y (ω) are the original (denoised) wavelet coefficients
at a particular level, I is the Indicator function (1 if true, 0 if
false), and T is some threshold limit. In both instances, if a
particular wavelet coefficient, yi, is less than T , then ωi = 0.
With hard thresholding, the remaining coefficients are unal-
tered; however, soft thresholding shrinks these coefficients by
the threshold limit.
There are many ways to estimate the value of T , including
VisuShrink, SURE Shrink, and Bayes Shrink (Chang et al.
2000). With TIDe, we implement the last technique because
it establishes a thresholding rule that is optimal in terms of
minimizing the expected RMS error in the denoised time se-
ries under flexible assumptions for the true time series sig-
nal (i.e., it minimizes the Bayes Risk for a squared-error
loss function). Bayes Shrink employs soft thresholding be-
cause its optimal estimator yields a smaller risk than hard-
thresholding's estimator. The optimal threshold value is deter-
mined as follows (see Chang et al. 2000). In some instances,
the noise variance, σ2, may be known a priori.
If this is
not the case, it is estimated from the robust median estima-
tor (Donoho & Johnstone 1994):
σ =
Median(Y1(y))
0.6745
,
(4)
where Y1(y) represents the wavelet coefficients, y, at the low-
est level (or finest scale) of decomposition. Next, we esti-
4
Stevenson et al.
mate the variance of Y (y) at a particular level j, assuming zero
mean, by:
σ2
j =
1
n
n
Xi=1
Y 2
j (yi).
(5)
where n is the number of wavelet coefficients at that level.
Our observation model (data = signal + noise) tells us that
+ σ2, where σ2
σ2
j = σ2
x is the signal variance. To account for
x
the case where σ2 > σ2
j , we calculate σx as follows:
σx =qmax(σ2
j
- σ2,0).
(6)
Finally, the optimal threshold at a particular level is deter-
mined by:
4. LIGHT-CURVE FITS AND RESULTS
We present the scaling of the RMS model residuals vs. bin
size (a test of correlation in time) in Figure 2 for all four 4.5-
µm Spitzer observations. Figure 3 displays our reanalysis of
GJ 436 data (Ballard et al. 2010a) plus three new Spitzer light
curves at 4.5 µm. Two fortuitous detections of UCF-1.01
appeared during atmospheric characterization observations of
GJ 436b (Stevenson et al. 2010). Using these data and a ten-
tative detection at 8.0 µm (see Section 4.6) to estimate its or-
bital period, we extrapolated UCF-1.01 transit times forward
by six months to predict an event (2011 July 30) during the
next observing window. We supply correlation plots and his-
tograms in Appendix A and the full set of best-fit parameters
from a 2.4×106-iteration joint fit in Appendix B. Below, we
discuss each observation in detail to explain how we arrived
at the final results.
Tj =
σ2
σx
.
(7)
10-2
2010 January 28
UCF-1.01 Ingress
UCF-1.01 Duration
S
M
R
10-3
10-4
10-5
10-1 100
101
102
103
2011 January 24
UCF-1.01 Ingress
UCF-1.01 Duration
10-3
10-4
S
M
R
2010 June 29
UCF-1.01 Ingress
UCF-1.01 Duration
100
101
102
103
2011 July 30
UCF-1.01 Ingress
UCF-1.01 Duration
10-3
10-4
10-5
10-3
10-4
In the event that σ2 > σ2
cients are set to zero.
j (Tj = ∞), all of the wavelet coeffi-
In this paper, we use the Biorthogonal 5.5 discrete wavelet
from the PyWavelets package to apply soft thresholding with
Bayes Shrink to the designated scale levels.
3.3. TIDe Example
We generate a series of 1000 test frames, each containing a
2D Gaussian with a width of 0.7 pixels, a peak flux of 1000,
and centered at (4.5, 4.5) in a 10×10 frame with the lower-left
corner indexed as (0, 0). We then added a background flux
offset of 100 and applied Poisson noise to each frame. We
performed 2D Gaussian centering to derive the blue points
plotted in Figure 1. For the points in red, we applied TIDe
to the frames using a Biorthogonal 5.5 discrete wavelet (from
the PyWavelets package) then recalculated the image centers
with the same 2D Gaussian centering routine. In each case,
only the y component of the position is plotted for each frame.
Using TIDe, the standard deviation in the pointing about the
true center decreased from 0.019 to 0.011 pixels, for a typical
improvement of ∼40%. We see even better results with the
2010 January 28 data set, where TIDe improved the pointing
precision by ∼70%.
0.06
0.04
0.02
0.00
None
TIDe
r
e
t
n
e
C
e
u
r
T
m
o
r
F
s
l
e
x
P
i
−0.02
−0.04
−0.06
0
200
400
Frame
600
800
1000
FIG. 1.- Illustrative example of TIDe that compares image centers
from simulated noisy frames (blue) to their denoised counterparts
(red). In each case, only the y component of the position is plotted
relative to the true center. In this typical example, the standard devi-
ation in the pointing (a measure of precision) decreased from 0.019
to 0.011 pixels.
10-5
100
102
101
Bin Size [sec]
103
10-5
100
102
101
Bin Size [sec]
103
FIG. 2.- Normalized RMS residual flux vs. bin size (in black) for four 4.5-
µm light curves. Black vertical lines at each bin size depict 1σ uncertainties
on the RMS residuals (RMS/√2N, where N is the number of bins). The red
lines show the predicted standard error for Gaussian noise. The dotted and
dashed lines indicate the scale length of UCF-1.01's best-fit transit ingress
and duration times, respectively. The excess RMS above the red line in the
top left panel indicates correlated noise at timescales near UCF-1.01's transit
duration and is discussed in Section 4.1.
4.1. 2010 January 28 (4.5-µm Spitzer Observation)
Spitzer program 541 (Sarah Ballard, P.I.) monitored GJ 436
continuously for ∼18 hours using 0.1-second exposures. The
short exposure time allows us to apply TIDe to the lowest
four levels (L4) of wavelet decomposition (see Section 3),
resulting in a maximum affected time resolution of 1.6 sec-
onds. In Stevenson et al. (2010), we detect Spitzer pointing
changes on timescales as short as ∼5 seconds, longer than
TIDe's timescale. In calculating image centers with and with-
out TIDe, we find that the position consistency between con-
secutive denoised frames improves by more than a factor of
three, resulting in an RMS of 0.0015 pixels in x and 0.0011
pixels in y. More precise image centers decrease flux scat-
ter with smaller aperture sizes and aid the BLISS map in
modeling the position-dependent systematics. We apply 5×-
interpolated aperture photometry to the unmodified frames to
avoid the computationally prohibitive calculation of estimat-
Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system
5
mained elevated until the end of the observation. We apply
a BLISS map bin size (x, y) of 0.007 × 0.005 pixels and set
the minimum number of points per bin to six to disregard the
observed excursion.
1.0005
1.0000
0.9995
d
l
x
F
d
e
z
i
l
a
m
r
o
N
0.9990
0.9985
0.9980
−0.20
d
c
c
c
c
b
l
x
u
F
d
e
z
i
l
a
m
r
o
N
d
c
Binned Data
Best Fit
1.0004
1.0002
1.0000
0.9998
0.9996
−0.2
−0.1
0.0
0.1
BJD - 2455225
0.2
0.3
0.4
0.5
FIG. 4.- Full light curve from 2010 January 28 Spitzer observa-
tion with transits of UCF-1.01 (right) and UCF-1.02 (left). The flux
values are corrected for systematics, normalized to the system bright-
ness, and binned (with 1σ error bars). The solid line depicts the best-
fit model. Excluding the excess flux near 2455225.23 in the model
fit does not significantly alter the best-fit solution.
4.2. 2010 June 29 (4.5-µm Spitzer Observation)
Our Spitzer program 60003 (Joseph Harrington, P.I.) mon-
itored GJ 436 for six hours using 0.4-second exposures. The
mean image center is located at pixel (15, 25), near the top
of the 32×32 array, thus restricting aperture sizes to ≤5.50
pixels. Using a background sky annulus from 10 to 30 pixels,
we find that the lowest SDNR occurs with a 5×-interpolated
aperture 5.00 pixels in radius. The BLISS map uses bins of
size 0.006 × 0.009 pixels and with at least four points per
bin. We test image centers generated from L3 TIDe (3.2-
second maximum time resolution) but find no improvement
in the SDNR. This is likely due to the smaller improvement
in image centers and significantly larger aperture size, rela-
tive to the 2010 January 28 dataset. The final analysis did not
use TIDe. For this dataset, the telescope pointing does not
stabilize until midway through the transit of UCF-1.02 (see
Figure 5). As a result, the position-dependent systematic is
poorly constrained during and prior to the transit. This may
be the source of UCF-1.02's variable transit duration, which
decreases with smaller aperture sizes. More observations are
necessary to confirm its parameters.
4.3. 2011 January 24 (4.5-µm Spitzer Observation)
Spitzer program 60003 performed a second six-hour ob-
servation of GJ 436 with 0.4-second exposures. We find
that 10×-interpolated aperture photometry outperforms 5×-
interpolated and minimizes SDNR with an aperture size of
5.25 pixels and a background sky annulus from 10 to 30 pix-
els. We flag 54 frames (28,426 – 28,479) as bad due to a
one-pixel horizontal shift, as observed previously in a dataset
above. We clip the first 6,000 observations due to a strong
increase in flux, possibly due to stellar activity (see Figure 3).
Near 2455585.771, we observe a sudden shift in the telescope
pointing that, again, correlates with an increase in background
noise. To remove this excursion from our models, the BLISS
map uses bins with eight or more points and a size of 0.016
× 0.008 pixels. As with the previous dataset and for the same
reasons, TIDe centers have little effect on the resulting photo-
metric light curve.
4.4. 2011 February 1 (3.6-µm Spitzer Observation)
2010 Jan ary 28
2010 J ne 29
2011 Jan ary 24
2011 J ly 30
−0.15
−0.10
−0.05
0.00
Time From Predicted Center of Transit [Days]
0.05
0.10
0.15
0.20
FIG. 3.- Four 4.5-µm Spitzer light curves of GJ 436 with best-fit mod-
els. The flux values are corrected for systematics, normalized to the system
brightness, and binned (with 1σ error bars). Light curves are vertically sep-
arated for ease of comparison. The single GJ 436b eclipse, four UCF-1.01
transits, and two UCF-1.02 candidate transits are indicated by the letters b, c,
and d, respectively. The transits distinguish themselves by their consistency
in depth and duration. Although UCF-1.01's 2010 January 28 best-fit tran-
sit time is 20 ± 7 minutes earlier than our predicted time (dashed line), the
parameter's probability distribution is bimodal (see Figure 14) and the other
peak is only 6 ± 7 minutes early. We quote the median transit time in Table 2
to encompass both possibilities. The three remaining UCF-1.01 transit times
(see Table 2) occur within five minutes of the predicted times and have a typ-
ical uncertainty of ±3 minutes. The episodic scatter in flux is most likely
due to stellar activity, which is expected for an M dwarf and seen in many
observations of this system. Using the non-detection of GJ 436b in the 2011
January 24 dataset, we place a 3σ upper limit of 95 ppm on its eclipse depth,
resulting in a brightness temperature <780 K. This new 4.5-µm secondary-
eclipse observation supports a methane-deficient and carbon monoxide-rich
dayside atmosphere (Stevenson et al. 2010).
ing uncertainties for the denoised frames, which are correlated
in time.
Photometry generates consistent transit depths for all tested
apertures from 1.25 to 4.50 pixels, but an aperture size of 2.25
pixels produces the lowest standard deviation of the normal-
ized residuals (SDNR). We estimate the background flux us-
ing an annulus from 7 to 15 pixels centered on the star. The
light curve (see Figure 4) exhibits a strong initial increase in
pixel sensitivity that we do not model (preclip, q = 10,000).
As with B10, we detect excess flux (possibly due to stellar ac-
tivity) near BJD 2455225.23 that contributes to the observed
correlated noise in Figure 2. We note that frames 19,780
– 19,839, 82,180 – 82,239, 165,380 – 165,439, 419,780 –
419,839, and 451,780 – 451,839 are shifted horizontally by
one pixel, so we flag these frames as bad. A probable micro-
meteor impact at BJD ∼2,455,224.976 caused a sudden shift
in pointing before returning to its original location. Simulta-
neously, the background scatter increased by ∼50% and re-
Stevenson et al.
l
x
u
F
d
e
z
i
l
a
m
r
o
N
1.0015
1.0010
1.0005
1.0000
0.9995
0.54
Binned Data
Best Fit
GJ 436b
0.62
0.56
Orbital Phase (2.64-day period)
0.58
0.60
6
x
n
i
n
o
i
t
i
s
o
P
l
e
x
P
i
15.05
15.00
14.95
14.90
14.85
14.80
26.0
y
n
i
n
o
i
t
i
s
o
P
l
e
x
P
i
25.8
25.6
25.4
25.2
25.0
y
n
i
n
o
i
t
i
s
o
P
l
e
x
P
i
25.8
25.7
25.6
25.5
25.4
25.3
25.2
UCF-1.01
UCF-1.02
0.55
0.60
0.65
BJD - 2455225
0.70
0.75
360
320
280
240
200
160
120
80
40
14.86 14.88 14.90 14.92 14.94 14.96 14.98 15.00
Pixel Position in x
FIG. 5.- Image centers vs.
time (upper 2 panels) and pointing his-
togram (lower panel, number of image centers within a given bin)
for the 2010 June 29 dataset. The times during transit are shaded in
gray. Initial telescope drift hampers our ability to effectively model
position-dependent systematics during and prior to the UCF-1.02
transit.
Our Spitzer program 60003 also observed GJ 436 at 3.6 µm
with 0.1-second exposures. We apply 5×-interpolated aper-
ture photometry with an aperture size of 2.75 pixels and a
background sky annulus from 7 to 15 pixels. We clip the
first 10,000 frames due to a steep ramp and frames 70,000
– 125,000 due to suspected stellar activity (see Figure 6). Be-
cause GJ 436b's time of secondary eclipse occurs during the
period of increased stellar activity, we do not fit the eclipse or
calculate uncertainties.
4.5. 2011 July 30 (4.5-µm Spitzer Observation)
Spitzer monitored GJ 436 for 4.3 hours using 0.4-second
exposures (program 70084, Joseph Harrington, P.I.). Pho-
tometry generates consistent transit depths for apertures be-
tween 1.75 and 6.00 pixels. The final run applies 10×-
interpolated, 5.00-pixel aperture photometry with a back-
ground sky annulus from 10 to 30 pixels. During these ob-
servations, the telescope pointing experiences two deviations,
at BJD 2,455,772.766 and 2,455,772.870. The background
variance increases with the first event but slightly decreases
with the second event. BLISS mapping utilizes a bin size of
0.012 × 0.006 pixels with a minimum of six points per bin
to exclude points from either excursion. Again, TIDe centers
FIG. 6.- Spitzer 3.6-µm light curve from 2011 February 1 with
GJ 436b's time of secondary eclipse shaded in gray. The flux val-
ues are corrected for systematics, normalized to the system bright-
ness, and binned (with 1σ error bars). The solid line depicts the
best-fit baseline model. Stellar activity prohibits us from fitting the
eclipse and measuring its depth. However, by visually comparing the
binned points within the shaded region to those immediately outside,
the data appear to be consistent with a relatively deep eclipse depth
(Stevenson et al. 2010).
have little effect on the resulting photometric light curve.
4.6. 2008 July 14 (8.0-µm Spitzer Observation)
Spitzer program 50056 (Heather Knutson, P.I.) observed GJ
436 for ∼70 hours from 2008 July 12 to 2008 July 15. At
the best aperture size of 3.75 pixels (and a background sky
annulus from 7 to 15 pixels), we find that the light curve ex-
hibits a measurable position-dependent systematic, identified
as pixelation (Stevenson et al. 2011). The BLISS map fits and
removes pixelation (see Figure 7) using a bin size of 0.022
× 0.022 pixels and at least four points per bin. We model
the initial time-dependent ramp using an asymptotically con-
stant exponential function after clipping the first 3,000 frames.
A sinusoidal function with a linear correction fits the phase
variation of GJ 436b (see Figure 8). We set a prior on the
inclination and semi-major axis of UCF-1.01 using the best-
fit results from the 4.5-µm joint fit. The UCF-1.01 transit at
BJD 2,454,662.328 is the same candidate transit reported by
Ballard et al. (2010a) using a ∼2.1-day period estimated from
EPOXI observations. Their Figure 5 incorrectly reports the
BJD. We used the timing of this transit to successfully predict
the 2001 July 30 transit. The best-fit radius ratio from both
transit events in this light curve is 0.010 ± 0.003, which is
consistent with the best-fit result using the four 4.5-µm light
curves.
4.7. Independent Analysis
We sought an independent analysis to confirm our results.
Nikole Lewis analyzed each of the 4.5-µm datasets without
knowing the times or depths of the transits.
In addition to
using her own photometric pipeline, she applied a new pixel-
mapping routine that shares a heritage with the method from
Ballard et al. (2010a). This new pixel-mapping method was
developed to recover the relative flux variations as a func-
tion of orbital phase from the Spitzer 3.6-µm and 4.5-µm
full orbit light curves of HD 189733b, HD 209458b, HAT-
P-2b, and HAT-P-7b (PI:Knutson; PID 60021). Similar to
the BLISS method, the pixel-mapping technique developed
by Lewis uses nearest neighbors to calculate flux as a func-
tion of position on the detector, but in her method the dis-
tances are weighted according to a Gaussian distribution. In
addition to stellar centroid positions, Lewis makes use of the
"noise pixel" parameter given in frame headers to determine
the nearest-neighbors to a given data point (Lewis et al., in
prep.). This routine improves on Ballard's method by calcu-
Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system
7
1.0015
1.0010
1.0005
1.0000
0.9995
0.9990
0.9985
0.9980
0.9975
6400
5600
4800
4000
3200
2400
1600
800
0
15.0
14.8
14.6
14.4
14.2
y
n
i
n
o
i
t
i
s
o
P
l
e
x
P
i
14.0
14.40
14.45
14.60
14.50
Pixel Position in x
14.55
14.65
14.70
15.0
14.8
14.6
14.4
14.2
y
n
i
n
o
i
t
i
s
o
P
l
e
x
P
i
14.0
14.40
14.45
14.50
14.60
Pixel Position in x
14.55
14.65
14.70
FIG. 7.- BLISS map (top) and pointing histogram (bottom) for the
2008 July 14 dataset. Pixelation, a position-dependent systematic, is
depicted by the colors in the BLISS map, where redder (bluer) col-
ors indicate more (less) flux within the aperture. Peaks repeat every
0.2 pixels because we applied 5×-interpolated aperture photometry.
Smaller interpolation factors result in larger spacing between peaks
but also a stronger systematic between peaks. The horizontal and
vertical black lines depict pixel boundaries.
lating the pixel map at each iteration without being computa-
tionally prohibitive.
Pixel mapping is essential to detecting the weak transit sig-
nals in these data. For example, the 2010 January 28 dataset
requires an accurate pixel mapping routine, at minimum, to
detect UCF-1.01 and benefits from more precise image cen-
ters with TIDe to more clearly distinguish UCF-1.02. We
have found that without a pixel-mapping routine, one can-
not reproduce all of the observed transits. Lewis uncovered
transits of UCF-1.01 in the 2010 June 29, 2011 January 24,
and 2011 July 30 datasets with ease and, once informed of
the additional planet, identified both UCF-1.02 transits and
the remaining UCF-1.01 transit (see Figure 9). Her final tran-
sit times, depths, and durations for both planets are all within
1.5σ of our best-fit results.
4.8. EPOXI Observation
NASA's EPOXI mission observed GJ 436 nearly continu-
ously during 2008 May 5 – 29 (Ballard et al. 2010b). The
light-curve data are available from EPOXI's archive. After
masking the transits of GJ 436b, we divide the light curve
1.0006
1.0004
1.0002
1.0000
0.9998
0.9996
0.9994
0.9992
b
c
b
c
b
Binned Data
Best Fit
l
x
u
F
d
e
z
i
l
a
m
r
o
N
0.5
1.0
1.5
2.0
BJD - 2454660
2.5
3.0
FIG. 8.- Spitzer light curve of GJ 436 at 8.0 µm. The flux values are
corrected for systematics, normalized to the median system bright-
ness, and binned (with 1σ error bars). The solid blue line depicts the
best-fit model. The light curve contains two eclipses (first and last
events) and one transit (center event) of GJ 436b and two transits
(second and fourth events) of UCF-1.01. The two UCF-1.01 tran-
sit depths have a combined significance of 2σ, which is insufficient
to claim a detection, but we used the timing of the latter to predict
the 2001 July 30 transit. The difference in brightness temperatures
between GJ 436b's dayside and nightside causes the observed sinu-
soidal variation in the light curve. The peak-to-peak flux difference
is 200 ± 50 ppm (4σ significance). This corresponds to a brightness
temperature difference of 110 ± 60 K, which favors a relatively ef-
ficient dayside-to-nightside energy redistribution. The peak flux is
shifted by 0.7 ± 4.6 hours prior to secondary eclipse.
by the median flux value, phase it according to the best-fit
UCF-1.01 period (see Table 2), and bin the results. Figure 10
illustrates a visible decrease in the observed flux at the correct
phase that is consistent with the transit depth and duration of
UCF-1.01 derived from the Spitzer data. The quality of the
light curve is such that the data neither prove nor disprove the
existence of UCF-1.01.
5. DISCUSSION
Without continuous monitoring of GJ 436 for two consecu-
tive transits at the most photometrically-precise wavelengths
(3.6 and 4.5 µm), we isolate the true period from integer mul-
tiples or whole number fractions by other means. Integer mul-
tiples (i.e., 2, 3, 4...) of the orbital period (see Table 2) cannot
account for all of the observed transits; whole number frac-
tions (i.e., 1/2, 1/3, 1/4...) are eliminated by investigating
the bevy of available GJ 436 Spitzer data at predicted tran-
sit times. We find evidence against periods of ∼0.6829 and
∼0.4553 days by non-detections of UCF-1.01 in a 2008 Jan-
uary 30 observation at 3.6 µm and a 2008 June 11 observation
at 8.0 µm, respectively. The single UCF-1.01 detection in the
2010 January 28, 18-hour observation dismisses even shorter
periods.
5.1. Eliminating False Positives
GJ 436's large proper motion (across the sky) enables us to
eliminate astrophysical false positives that could mimic the
observed periodic decrease in flux. Over our 1.5-year ob-
servational baseline of 4.5-µm detections, the system moves
∼1.8′′, equivalent to 1.5 pixels in Spitzer's InfraRed Array
Camera. With aperture sizes as small as 1.25 pixels for the
first (2010 January 28) and last (2011 July 30) observations,
we find that the transit signals from UCF-1.01 are clearly dis-
tinguished against the background noise. This limits the lo-
cation of a potential background source (such as an eclipsing-
binary star system, Torres et al. 2011) to the overlapping re-
gion within both apertures. Using observations from the Very
Large Telescope (VLT, see Figure 11, Rousset et al. 2003;
Lenzen et al. 2003) and Canada France Hawaii Telescope
(CFHT, see Figure 12, Rigaut et al. 1998) with adaptive op-
tics imaging instruments at two different epochs, we elimi-
8
Stevenson et al.
TRANSIT MODEL BEST-FIT VALUES AND OTHER PARAMETERS
TABLE 2
UCF-1.01
UCF-1.02
Parameters
Rp/R⋆
i [◦]
a/R⋆
Impact Parameter
Transit depth [ppm]
Duration [t4-1, hr]
Ingress/Egress [hr]
Transit Times [MJDT DB]b
Mean Period [Days]
Ephemeris [MJDT DB]b
Radius [R⊕]
Mass [M⊕]d
a
0.0138 ± 0.0009a
85.17+0.8
- 0.16
9.10 ± 0.07a
0.77+0.02
- 0.15
190 ± 25
0.76+0.15
- 0.03
0.025+0.002
- 0.004
5225.090+0.004
c
- 0.005
5376.7078+0.0014
- 0.0021
5585.6889+0.0020
- 0.0018
5772.8069+0.0009
- 0.0029
1.365862 ± 8×10- 6
5772.8086 ± 0.0016
a
0.66 ± 0.04
0.28 ± 0.06
a
a
0.0136 ± 0.0012
–
–
–
186 ± 30a
1.04+0.26
- 0.15
<0.1a
a
5225.026 ± 0.003a
5376.568+0.003
- 0.007
a
–
–
–
–
0.65 ± 0.06
0.27 ± 0.07
a Fitted values.
b MJD = BJD - 2,450,000.
c We choose the median value because the distribution is bimodal.
d Assuming an Earth-like density of 5.515 g/cm3.
FIG. 9.- Four 4.5-µm Spitzer light curves of GJ 436 with best-
fit models from an independent analysis by Lewis. She corrects
flux measurements for intrapixel sensitivity variations using a pixel-
mapping technique and for the presence of the well documented
Spitzer time-dependent systematic (ramp). A fixed-width symmet-
ric Gaussian fits centroid positions in the region near the bright-
est pixel in each subarray frame. The best photometric aperture
size is 2.25 pixels for the 2010 January 28 dataset and 5.0 pixels
for the other datasets. The non-linear limb-darkening coefficients
for GJ 436 are those from Knutson et al. (2011). After the loca-
tion of the transit(s) in each dataset were identified individually,
Lewis performed a simultaneous fit between all four datasets using
a Levenberg-Marquardt minimization scheme. A MCMC algorithm
determined the uncertainty in the fit parameters as well as identified
other possible solutions. The goal of this analysis was to confirm
the presence and shape of transit(s) in each dataset. Improvements
to treatment of the systematics in these observations is possible, but
they are unlikely to significantly change the estimated planetary pa-
rameters.
1''
2011 July 30
2011 January 24
2010 June 29
2010 January 28
2008 July 14
2007 March 20
N
E
l
x
u
F
d
e
z
i
l
a
m
r
o
N
1.0002
1.0001
1.0000
0.9999
0.9998
0.9997
Binned EPOXI Data
UCF-1.01 Transit Depth
−0.4
−0.2
0.0
Phase
0.2
0.4
FIG. 10.- EPOXI light curve phased to the period of UCF-1.01
using the best-fit period and nearest ephemeris time (2008 July 14
dataset). Blue circles represent the binned EPOXI data with 1σ un-
certainties. The red cross depicts the duration and depth (with a 1σ
uncertainty) of UCF-1.01's transit. The EPOXI data are consistent
with a UCF-1.01 transit.
nate background stars up to 12.7 and 9.3 magnitudes fainter,
respectively, than GJ 436 at a confidence of 5σ.
FIG. 11.- Very Large Telescope H-band observation on 2007 March 20
using the NAOS-CONICA instrument with adaptive optics (Montagnier et al.
2012). We search for faint background systems by blocking the light from GJ
436 using a 0.7′′ Lyot coronagraphic mask. The dark green lines are mask
support wires. The "+" symbols indicate the position of the GJ 436 system for
this observation and at each transit epoch of UCF-1.01. Red circles indicate
the minimum photometric aperture size (1.25 pixels) for which transit signals
from the first and last confirmed events may still be clearly distinguished
against the background noise. If a background system were the source of the
transit-like events, it must put light in the overlapping region. To produce
the observed transit depth, the hypothetical system must be no more than
9.3 magnitudes fainter than GJ 436, assuming a total eclipse of one of the
objects. We eliminate objects brighter than ∆H = 12.7 relative to GJ 436 with
a confidence of 5σ. There are also no objects listed in any catalog within this
region.
5.2. Instability Hypothesis
In this section we consider an alternate hypothesis to that
of detecting two sub-Earth-sized exoplanets, that the observed
variations are the result of instrumental or stellar instabilities.
Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system
9
2011 July 30
2011 January 24
2010 June 29
2010 January 28
2008 July 14
2''
2000 April 19
N
E
FIG. 12.- Canada France Hawaii Telescope (CFHT) K-band observation
obtained 2000 April 19 using the adaptive optics bonnette (PUEO). The "+"
symbols indicate the position of the GJ 436 system for this observation and at
each transit epoch of UCF-1.01. Red circles indicate the minimum photomet-
ric aperture size (1.25 pixels) for which transit signals from the first and last
confirmed events may still be clearly distinguished against the background
noise. We eliminate background objects within the overlapping region with
∆K = 9.3 at a 5σ confidence limit.
We begin by calculating the probability of observing UCF-
1.01 and UCF-1.02 by chance. Then, we quantify the occur-
rence rate of transit-like instabilities and estimate the proba-
bility that these instabilities are periodic. Finally, we compare
our model fits to the null hypothesis, which is expressed by
a model that does not contain planet parameters, to see if the
additional free parameters are justified.
In search of transit signals in the GJ 436 system, we ex-
amined 11 light curves at 3.6 and 4.5 µm (not counting the
2011 July 30 dataset in which we predicted the transit). Both
channels have the photometric stability necessary to detect
GJ 436c. Of the 71.3 hours of data, there are eight transit
or eclipse events of GJ 436b, each lasting ∼1-hour in dura-
tion. Since we cannot distinguish overlapping transits, we
have 63.3 hours of usable data with an average light-curve
duration of ∼5.75 hours. Given that the period of UCF-1.01
is 1.3659 days, the probability that a transit will occur in a
typical event is 17.5%. Using the binomial distribution, we
calculate a 30.1% chance of observing three or more transits
of UCF-1.01 in the 11 available light curves. Recall that our
fourth transit event of UCF-1.01 was predicted, rather than
occurring by chance, so it does not enter into the calculation.
We repeat the above calculation for UCF-1.02 but must first
estimate its orbital period by considering the transit duration
ratio between itself and GJ 436b. We find that the durations
are nearly identical; however, both planets are unlikely to oc-
cupy the same orbit. So, we perform two sets of calculations:
one for each side of the 1-sigma uncertainty in UCF-1.02's
transit duration. Using a period of 5.563 days, the probability
of observing two or more UCF-1.02 transits is 7.9%. Using
a period of 1.785 days, the probability of observing two or
more transits increases to 44.6%. The combined probability
of observing both planets ranges from 2.3% to 13.4%.
We compare these results to the alternate hypothesis,
namely that the observed variations are the result of instru-
mental or stellar instabilities. To begin, we analyzed nearly
120 hours of GJ 1214 data at 4.5 µm (Spitzer program 70049).
This M dwarf is similar to GJ 436 and should exhibit similar
levels of activity (stellar instabilities). If the instabilities are
instrumental, then it should not matter which star we analyze
unless the instabilities are flux-dependent (GJ 436 is almost 3
magnitudes brighter than GJ 1214 in the infrared). From our
GJ 1214 light-curve results, we identify two transit-like events
based on their depths (>200 ppm) and durations (∼1 hour).
Assuming these events are not the result of planet transits, for
any given hour of 4.5-µm observations, there is a 1.7% prob-
ability of having an instability event. We then apply the bino-
mial distribution to determine that the probability of detecting
five or more instabilities in 63.3 hours of data is 0.42%. Re-
call that we do not count the 2011 July 30 dataset or use times
during GJ 436b transits/eclipses. If we assume that the insta-
bilities only appear at 4.5 microns, the probability of detect-
ing five or more instabilities in 44.7 hours of data decreases
to 0.088%.
Since we cannot find a physical mechanism for reducing the
stellar flux in a transit-like way with a repeatable period, any
observed instabilities must be random events. We consider
the probability of detecting four random instability events that
happen to coincide with a given period (in this case, 1.3659
days). The first two instability events establish the "period"
under consideration. As calculated above, the third and fourth
instability events each have a 1.7% probability of occurring
within 30 minutes of the established period (total time win-
dow is 1 hour). Their combined probability is 0.029%.
We conclude that the single-planet hypothesis is 72 times
more likely than the most favorable instrumental/stellar-
instability scenario and 1040 times more likely than detect-
ing four random instability events that happen to coincide
with a given period. The two-planet hypothesis is 5.5 – 32
times more likely than the most favorable instrumental/stellar-
instability scenario and 79 – 460 times more likely than de-
tecting four random instability events that happen to coincide
with a given period.
Finally, we test the strength of our two sub-Earth-sized exo-
planet candidates by comparing various fits to the null hypoth-
esis, which asserts that there are no new planets. Recall that
a lower ∆BIC value indicates that the additional free parame-
ters are warranted in the model fit. Relative to the null hypoth-
esis, ∆BIC decreases by 11.4 when we include UCF-1.01's
transit parameters in a joint model fit. Alternatively, if we
add only UCF-1.02's transit parameters then ∆BIC increases
by 36.2. Including both planets' transit parameters in a joint
model fit results in an increase in ∆BIC of 14.5, relative to the
null hypothesis. Thus, the BIC favors a model that includes
UCF-1.01 but disfavors models that include UCF-1.02. This
result is directly related to the number of observed transits
for each planet candidate and indicates that we need to obtain
more than two transit observations of UCF-1.02 to increase
the detection significance and surpass the BIC's penalty for
additional free parameters. We conclude that the available
data support UCF-1.01 as a strong exoplanet candidate and
signify that UCF-1.02 is a weak exoplanet candidate.
5.3. Radial-Velocity Constraints
The 3.6-meter ESO telescope at La Silla Observatory
(Mayor et al. 2003; Pepe et al. 2004) utilized the HARPS
spectrograph with the settings described by Bonfils et al.
10
Stevenson et al.
(2011) to obtain 171 observations of GJ 436 at 550 nm.
Xavier Bonfils (personal communication) provided us with
the extracted, unpublished RV measurements so that we
could attempt to constrain the mass of UCF-1.01. We re-
tained 159 data points (12 were removed due to the Rossiter-
McLaughlin effect). To these data, we added 41 GJ 436b
primary transit times (Knutson et al. 2011, and references
therein), 14 GJ 436b secondary eclipse times (Stevenson et al.
2010; Knutson et al. 2011), and an 8.0-µm photometric light
curve from Deming et al. (2007). The light curve (retrieved
from the Infrared Processing and Analysis Center, IPAC) is
binned into 445 points and normalized to remove the time-
dependent ramp.
We apply a two-planet model with the transit ephemeris for
the second planet fixed to the best-fit value listed in Table 2
and the eccentricity fixed to 0. The fit utilizes the empirical
stellar density calibration of Enoch et al. (2010) to determine
the stellar mass, in addition to other system parameters, grant-
ing this fit a much broader scope than the modeling described
by Campo et al. (2011). We employ a Levenberg-Marquardt
minimizer to find the best-fit parameters to our model and a
Markov-chain Monte Carlo routine with 106 iterations to esti-
mate uncertainties. We express our χ2 function as follows:
t j
χ2 =Xi (cid:20) vi - vi
σv,i (cid:21)2
(cid:20) ok - ok
σo,k (cid:21)2
+ Xk
+Xj (cid:20)t j -
σt, j (cid:21)2
+Xm (cid:20) pm - pm
σp,m (cid:21)2
(8)
0.50
0.45
0.40
]
⊕
UCF-1.01
UCF-1.02
KOI-961.03
M
[
s
s
a
M
t
e
n
a
P
l
0.35
0.30
0.25
0.20
0.15
0.10
3
12
]
2
s
/
m
[
y
t
i
v
a
r
G
e
c
a
f
r
u
S
t
e
n
a
P
l
10
8
6
4
2
3
Mars
4
UCF-1.01
UCF-1.02
KOI-961.03
5
6
7
8
Earth
Venus
Mars
Mercury
4
5
6
7
8
Planet Density [g/cm3 ]
where v, t, o, and p represent the HARPS radial velocities,
primary-transit times, secondary-eclipse times, and photomet-
ric data, respectively. The over-lined quantities indicate com-
puted values and σ represents the uncertainty for each mea-
surement. We adjust for transit-eclipse light travel times and
for leap seconds in this fit. Using the above data with an es-
timated stellar jitter of 1.7 m/s, we do not detect the signal of
the second planet but cannot repudiate its existence. The 3σ
upper limit of the semi-amplitude is 0.6 m/s, corresponding
to an upper limit of 0.6 M⊕, which is larger than our mass
constraints using the density arguments below.
5.4. Mass Constraints
Unable to effectively constrain the mass of UCF-1.01 using
RV data, we consider a range of possible bulk densities for
a terrestrial-sized planet (see Figure 13). Given a mean bulk
density between 3 and 8 g cm- 3, we limit the mass of UCF-
1.01 to be 0.15 – 0.40 M⊕ and estimate the surface gravity
to be 0.36 – 0.94 g. We place similar limits on the mass and
surface gravity of UCF-1.02. Assuming an Earth-like density
of 5.515 g cm- 3, we estimate masses of 0.28 and 0.27 M⊕ for
UCF-1.01 and UCF-1.02, respectively, which correspond to
surface gravities of ∼0.65 times that on Earth.
5.5. Orbital Constraints
UCF-1.01 may exhibit TTVs due to gravitational interac-
tions with GJ 436b in a near-2:1 orbital resonance or with
UCF-1.02, which has an unknown orbit. This may explain
why UCF-1.01's transit time is 20 minutes early in the 2010
January 28 data set; however, the parameter's probability dis-
tribution is bimodal (see Figure 14) and the other peak is only
6 ± 7 minutes early. The three remaining UCF-1.01 transit
FIG. 13.- Mass and surface-gravity constraints on UCF-1.01 (solid lines)
and UCF-1.02 (dashed lines). Bold lines depict the best-fit values and thin
lines depict the upper and lower 1σ uncertainties. Exoplanet KOI-961.03 and
solar-system planets are included for reference.
times occur within five minutes of their predicted times and
have a typical uncertainty of ±3 minutes. More precise ob-
servations could establish whether these deviations are TTVs.
Using the known orbital parameters of GJ 436b and UCF-
1.01, we performed orbital-stability simulations using the
Mercury numerical integrator (Chambers 1999). Assuming an
Earth-like density of 5.5 g cm-3, the predicted mass of UCF-
1.01 is 0.28 M⊕. We supplied the code with initial starting
conditions, listed in Table 3, based on transit times from the
2011 January 24 dataset. Our results indicate that the orbits
are stable out to at least 100 Myr. The best-fit line shows a
change in semi-major axis of 5.3e-06 au/Gyr. The osculat-
ing UCF-1.01 orbital parameters exhibit a periodic trend ev-
ery ∼35 years wherein the eccentricity varies between 0 and
0.21, the peak-to-trough inclination amplitude is 3.2◦, and
TTVs vary from ±200 to ±3 minutes. A ∼40-day periodic
trend is also evident but with smaller variations in the oscu-
lating orbital parameters. Due to UCF-1.01's relatively small
mass, variations in GJ 436b's orbital parameters over the 35-
year timespan are below Spitzer's sensitivity limits. Next-
generation facilities may be able to constrain UCF-1.01's or-
bital parameters through improved RV measurements or by
measuring its time of secondary eclipse.
5.6. Atmospheric Constraints
UCF-1.01 is unlikely to have retained any original atmo-
sphere due to its weak gravitational field, close proximity to
Two nearby sub-Earth-sized exoplanet candidates in the GJ 436 system
11
INITIAL ORBITAL PARAMETERS FOR GJ 436B AND UCF-1.01
TABLE 3
Parameter
Semi-major Axis
Eccentricity
Inclination
Argument of Periapsis
Longitude of the Ascending Node
Mean Anomaly
GJ 436b
0.0287 au
UCF-1.01
0.0185 au
0.1371
86.699◦
351.0◦
0.0◦
282.6◦
0
85.1◦
0.0◦
0.0◦
90◦
its host star, and estimated 6-Gyr age of the system (Torres
2007). The planet receives a substantial soft x-ray and ex-
treme ultraviolet (XUV) flux; we estimate 700 – 900 erg
cm- 2 s- 1 (Sanz-Forcada et al. 2011; Ehrenreich et al. 2011),
or ∼1,000 times the present XUV flux received by the Earth.
Such an intense XUV flux leads to a very hot thermosphere
and subsequent hydrodynamic escape (Tian 2009). Shortly
after formation, outgassing from an Earth-like, silicate-rich
mantle could have produced an initial water-vapor-rich atmo-
sphere for UCF-1.01 (Schaefer et al. 2011). However, the wa-
ter vapor would readily have been photolyzed by ultraviolet
radiation at high altitudes, leading to a hydrogen-dominated
thermosphere that likely extended to the planet's Roche dis-
tance of ∼25,000 km (Erkaev et al. 2007), given the planet's
low gravity. In this situation, the mass-loss rate for energy-
limited hydrodynamic flow (Erkaev et al. 2007) implies a hy-
drogen loss rate of about 8 × 1010 g s-1 (assuming an XUV
heating efficiency of 1), or 1.4 times the planet's mass lost
in 1 Gyr. This indicates that hydrogen was lost from UCF-
1.01's atmosphere very early in its history. Some heavy ele-
ments would have been entrained in the hydrodynamic flow,
but the early atmosphere would have become increasingly ox-
idized as hydrogen was lost. Carbon dioxide could then have
dominated at some later point in the atmosphere's history, but
even a CO2-rich atmosphere would be unstable. Scaling from
hydrodynamic models (Tian 2009), we estimate that carbon
would be lost from a CO2-rich atmosphere at ∼ 1 × 108 g
s-1, or 1% of the planet's mass over its lifetime – an amount
likely greater than the planet's initial inventory of CO2. At-
mospheres dominated by molecular nitrogen or oxygen would
be lost on even shorter timescales (Tian 2009).
UCF-1.01 could support a transient, present-day atmo-
sphere if recent impacts were to deliver volatiles rather than
preferentially erode any atmosphere, or if tidal heating were
to supply volatiles from the crust/mantle. The latter sce-
nario is particularly attractive if a recycling mechanism ex-
ists for any heavy atmospheric constituents (e.g., volcanic
emission of sulfur dioxide, followed by photolysis to sulfur
and oxygen atoms, dayside-to-nightside transport, condensa-
tion, and subsequent melting and re-vaporization of sulfur de-
posits). In this speculative scenario, UCF-1.01 could resem-
ble a hot Io that has lost its lighter and more volatile elements.
Any transient atmosphere will likely have a low surface pres-
sure and be highly extended, which could fill the Roche lobe
and/or produce a tail. Transit observations at ultraviolet wave-
lengths could confirm or rule out such an extended atmo-
sphere, and one might search particularly at wavelengths in
which atomic and ionized sulfur and oxygen would be ex-
pected to absorb. Given that volcanically supplied sodium
and potassium might be transient atmospheric constituents,
visible-wavelength transit observations might also prove use-
ful.
6. CONCLUSIONS
In this paper, we announced the detection of UCF-1.01
and UCF-1.02, two sub-Earth-sized transiting exoplanet can-
didates orbiting the nearby M dwarf GJ 436. Their detec-
tions were possible with BLISS mapping and Time-series Im-
age Denoising (TIDe), the latter of which is a novel wavelet-
based technique that decreases high-frequency noise in short-
cadence, time-series images to improve image centering pre-
cision. We presented four transits of UCF-1.01 and two tran-
sits of UCF-1.02 at 4.5 µm, an independent analysis that con-
firms our best-fit results within 1.5σ, an 8.0-µm phase curve
of GJ 436b that includes transits of UCF-1.01, and EPOXI
data that are consistent with the presence of a sub-Earth-sized
exoplanet. To definitively establish UCF-1.01 as a planet (to
be called GJ 436c), we require only a few hours of addi-
tional observations, preferably from another telescope or at
least at a different wavelength. Establishing UCF-1.02 as a
planet (to be called GJ 436d) would likely require an extended
observing campaign to constrain its period then successfully
predict a transit. Finally, we confirmed the GJ 436b 4.5-µm
results presented by Stevenson et al. (2010) through an addi-
tional non-detection during secondary eclipse; however, we
were unable to confirm the strong eclipse depth at 3.6 µm due
to stellar activity. The current data still support a methane-
deficient and carbon monoxide-rich dayside atmosphere.
We thank Tom Loredo for providing comments on TIDe
and members of the HARPS team for discussions and
for providing unpublished RV data with an independent
analysis. We acknowledge Drake Deming for supplying
the GJ 1214 data and leading the EPOXI mission; we also
acknowledge Heather Knutson, the PI of Spitzer program
50056. We thank contributors to SciPy, Matplotlib, and the
Python Programming Language, the free and open-source
community, the NASA Astrophysics Data System, and the
JPL Solar System Dynamics group for software and services.
This work is based on observations made with the Spitzer
Space Telescope, which is operated by the Jet Propulsion
Laboratory, California Institute of Technology under a
contract with NASA. Support for this work was provided by
NASA through an award issued by JPL/Caltech.
REFERENCES
Alonso, R. et al. 2008, Astron. Astrophys., 487, L5
Ballard, S., Charbonneau, D., Deming, D., Knutson, H. A., Christiansen,
J. L., Holman, M. J., Fabrycky, D., Seager, S., & A'Hearn, M. F. 2010a,
PASP, 122, 1341
Ballard, S., Christiansen, J. L., Charbonneau, D., Deming, D., Holman,
M. J., Fabrycky, D., A'Hearn, M. F., Wellnitz, D. D., Barry, R. K.,
Kuchner, M. J., Livengood, T. A., Hewagama, T., Sunshine, J. M.,
Hampton, D. L., Lisse, C. M., Seager, S., & Veverka, J. F. 2010b, ApJ,
716, 1047
Bean, J. L. & Seifahrt, A. 2008, A&A, 487, L25
Beaulieu, J. P., Carey, S., Ribas, I., & Tinetti, G. 2008, ApJ, 677, 1343
Bonfils, X., Delfosse, X., Udry, S., Forveille, T., Mayor, M., Perrier, C.,
Bouchy, F., Gillon, M., Lovis, C., Pepe, F., Queloz, D., Santos, N. C.,
Ségransan, D., & Bertaux, J.-L. 2011, ArXiv e-prints
Butler, R. P. et al. 2004, ApJ, 617, 580
Campo, C. J., Harrington, J., Hardy, R. A., Stevenson, K. B., Nymeyer, S.,
Ragozzine, D., Lust, N. B., Anderson, D. R., Collier-Cameron, A., Blecic,
J., Britt, C. B. T., Bowman, W. C., Wheatley, P. J., Loredo, T. J., Deming,
D., Hebb, L., Hellier, C., Maxted, P. F. L., Pollaco, D., & West, R. G.
2011, ApJ, 727, 125
Chambers, J. E. 1999, MNRAS, 304, 793
12
Stevenson et al.
Chang, S. G., Yu, B., & Vetterli, M. 2000, IEEE TRANSACTIONS ON
IMAGE PROCESSING, 9, 1532
Charpinet, S., Fontaine, G., Brassard, P., Green, E. M., van Grootel, V.,
Randall, S. K., Silvotti, R., Baran, A. S., Østensen, R. H., Kawaler, S. D.,
& Telting, J. H. 2011, Nature, 480, 496
Claret, A. 2000, A&A, 363, 1081
Deming, D. et al. 2007, ApJ, 667, L199
Demory, B.-O., Gillon, M., Waelkens, C., Queloz, D., & Udry, S. 2009, in
IAU Symposium, Vol. 253, IAU Symposium, 424–427
Donoho, D. L. & Johnstone, I. M. 1994, Biometrika, 81, 425
Ehrenreich, D., Lecavelier Des Etangs, A., & Delfosse, X. 2011, A&A, 529,
A80
A33
Montagnier, G., Bonfils, X., Ségransan, D., Udry, S., Beust, H., Beuzit, J.-L.,
Forveille, T., Delfosse, X., Mayor, M., & Santos, N. 2012, A&A, in prep.
Muirhead, P. S., Johnson, J. A., Apps, K., Carter, J. A., Morton, T. D.,
Fabrycky, D. C., Pineda, J. S., Bottom, M., Rojas-Ayala, B., Schlawin, E.,
Hamren, K., Covey, K. R., Crepp, J. R., Stassun, K. G., Pepper, J., Hebb,
L., Kirby, E. N., Howard, A. W., Isaacson, H. T., Marcy, G. W., Levitan,
D., Diaz-Santos, T., Armus, L., & Lloyd, J. P. 2012, ApJ, 747, 144
Pepe, F., Mayor, M., Queloz, D., Benz, W., Bonfils, X., Bouchy, F., Lo
Curto, G., Lovis, C., Mégevand, D., Moutou, C., Naef, D., Rupprecht, G.,
Santos, N. C., Sivan, J.-P., Sosnowska, D., & Udry, S. 2004, A&A, 423,
385
Pont, F., Gilliland, R. L., Knutson, H., Holman, M., & Charbonneau, D.
Ribas, I., Font-Ribera, A., & Beaulieu, J.-P. 2008, ApJ, 677, L59
Rigaut, F., Salmon, D., Arsenault, R., Thomas, J., Lai, O., Rouan, D., Véran,
J. P., Gigan, P., Crampton, D., Fletcher, J. M., Stilburn, J., Boyer, C., &
Jagourel, P. 1998, PASP, 110, 152
Rousset, G., Lacombe, F., Puget, P., Hubin, N. N., Gendron, E., Fusco, T.,
Arsenault, R., Charton, J., Feautrier, P., Gigan, P., Kern, P. Y., Lagrange,
A., Madec, P., Mouillet, D., Rabaud, D., Rabou, P., Stadler, E., & Zins, G.
2003, in Presented at the Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference, Vol. 4839, Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series, ed.
P. L. Wizinowich & D. Bonaccini, 140–149
Sanz-Forcada, J., Micela, G., Ribas, I., Pollock, A. M. T., Eiroa, C., Velasco,
A., Solano, E., & García-Álvarez, D. 2011, A&A, 532, A6
Schaefer, L., Lodders, K., Fegley, B., & Jr. 2011, ArXiv e-prints
Stevenson, K. B., Harrington, J., Fortney, J. J., Loredo, T. J., Hardy, R. A.,
Nymeyer, S., Bowman, W. C., Cubillos, P., Bowman, M. O., & Hardin,
M. 2011, ArXiv e-prints
Stevenson, K. B., Harrington, J., Nymeyer, S., Madhusudhan, N., Seager, S.,
Bowman, W. C., Hardy, R. A., Deming, D., Rauscher, E., & Lust, N. B.
2010, Nature, 464, 1161
Tian, F. 2009, ApJ, 703, 905
Torrence, C. & Compo, G. P. 1998, Bulletin of the American Meteorological
Society, 79, 61
Torres, G. 2007, ApJ, 671, L65
Tor |
1709.09895 | 1 | 1709 | 2017-09-28T11:19:32 | Solar Radiation Pressure Resonances in Low Earth Orbits | [
"astro-ph.EP"
] | The aim of this work is to highlight the crucial role that orbital resonances associated with solar radiation pressure can have in Low Earth Orbit. We review the corresponding literature, and provide an analytical tool to estimate the maximum eccentricity which can be achieved for well-defined initial conditions. We then compare the results obtained with the simplified model with the results obtained with a more comprehensive dynamical model. The analysis has important implications both from a theoretical point of view, because it shows that the role of some resonances was underestimated in the past, but also from a practical point of view in the perspective of passive deorbiting solutions for satellites at the end-of-life. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- 9 (2017)
Preprint 6 August 2018
Compiled using MNRAS LATEX style file v3.0
Solar Radiation Pressure Resonances in Low Earth Orbits
Elisa Maria Alessi,1(cid:63) G. Schettino,1 A. Rossi1 and G. B. Valsecchi1,2
1IFAC-CNR, via Madonna del Piano 10, 50019 Sesto Fiorentino (FI), Italy
2IAPS-INAF, via Fosso del Cavaliere 100, 00133 Rome, Italy
Accepted XXX. Received YYY; in original form ZZZ
ABSTRACT
The aim of this work is to highlight the crucial role that orbital resonances as-
sociated with solar radiation pressure can have in Low Earth Orbit. We review the
corresponding literature, and provide an analytical tool to estimate the maximum ec-
centricity which can be achieved for well-defined initial conditions. We then compare
the results obtained with the simplified model with the results obtained with a more
comprehensive dynamical model. The analysis has important implications both from
a theoretical point of view, because it shows that the role of some resonances was
underestimated in the past, but also from a practical point of view in the perspective
of passive deorbiting solutions for satellites at the end-of-life.
Key words: Solar Radiation Pressure -- Resonances -- Low Earth Orbits
1 INTRODUCTION
It is known that the effect of the solar radiation pressure
(SRP) on the orbital motion is a long-period variation in
eccentricity e and inclination i, along with one in longitude
of ascending node Ω and argument of pericenter ω, and that
the magnitude of this variation depends on the area-to-mass
ratio of the body (e.g., Beutler 2005). If we assume a per-
turbing potential which is averaged over the orbital period
of the body, the SRP effect can be subject to resonances,
associated with a commensurability among the rate of pre-
cession of the ascending node, the one of the argument of
pericenter and the mean apparent motion of the Sun, say nS.
As well-known, when a resonance occurs, the period of the
variation can become so large to give rise to quasi-secular
effects in the corresponding orbital element.
Motivated by the numerical results (Alessi et al.
2017a,b) obtained recently within the ReDSHIFT H2020
project (Rossi et al. 2017), the aim of this work is to de-
scribe the role of SRP resonances in the Low Earth Orbit
(LEO) region, showing that their contribution can provide
relatively significant eccentricity variations.
In the past, the effect of the solar radiation pressure
for orbits around the Earth was considered mainly in the
perspective of bodies characterised by a very high area-to-
mass ratio. We can mention works focused on the behaviour
of Geostationary Earth Orbits (GEO) (e.g., Valk et al. 2007;
Rosengren & Scheeres 2013; Casanova et al. 2015), or on
mission concepts aiming to exploit a solar sail to deorbiting
non-operational objects from Medium Earth Orbits (MEO)
(cid:63) E-mail: [email protected] (EMA)
© 2017 The Authors
(e.g., Lucking et al. 2012), or on the design of frozen orbits
for small-size objects (e.g., Colombo et al. 2012; Lantukh et
al. 2015). Colombo et al. (2012), in particular, analysed SRP
effects on high area-to-mass objects in order to design frozen
orbits for a swarm of small spacecraft dedicated to Earth
observation and telecommunication purposes. Starting from
the work by Krivov et al. (1996), they studied the behaviour
in eccentricity corresponding to (cid:219)Ω+ (cid:219)ω−nS = 0, by computing
the equilibrium points, and their stability, in a dynamical
system including the oblateness of the Earth and SRP.
From a theoretical point of view, the first works on the
resonances induced by SRP started from the analysis of the
resonances induced by the solar gravitational attraction. As
a matter of fact, the two disturbing potentials differ in the
first order of the expansion and in the amplitude of the per-
turbation (Hughes 1977). If we focus on the resonances af-
fecting the eccentricity evolution, Musen (1960) and Cook
(1962) were the first to highlight the existence of six SRP
resonances, and to show their location in the (i, a) plane.
Cook (1962), in particular, noted that, contrary to luniso-
lar gravitational resonances, the SRP resonances are able to
give a variation in eccentricity also in the case of circular
orbits. Musen (1960) labelled (cid:219)Ω + (cid:219)ω − nS as the 'most inter-
esting resonance', as done by all the following authors. In the
case of the gravitational perturbation, this resonance is also
known as evection resonance1 (e.g., Brouwer & Clemence
1961; Touma & Widsom 1998; Frouard et al. 2010). Bre-
iter (1999), basing his analysis on the work done by Cook
(1962), noted that 'resonant solar perturbations can be much
1 To be precise, 2( (cid:219)Ω + (cid:219)ω − nS).
E. M. Alessi et al.
2
stronger than the lunar ones'. He also considered (cid:219)Ω+ (cid:219)ω−nS as
the dominant resonance for prograde orbits and (cid:219)Ω − (cid:219)ω + nS
for retrograde orbits. In his work, the strength of a given
resonance is determined by the magnitude of the libration
region. As already mentioned, a fundamental contribution to
the topic was given by Hughes (Hughes 1977, 1980, 1981),
who tried to provide a more general treatment to the prob-
lem, by analysing also the effect due to high-order terms2.
He stressed that SRP resonances give rise to variations not
only in eccentricity, but also in inclination, longitude of the
ascending node and argument of pericenter. He considered
a resonance as dominant according to the magnitude of the
corresponding amplitude, function of semi-major axis a and
eccentricity. More recently, Celletti et al. (2017) recognised
the occurrence of semi-secular resonances, as the SRP ones,
in the LEO region, but they stated that 'Since the semi-
secular resonances occur at specific altitudes, the width of
such resonances is small, and since the air drag provokes
a decay of the orbits on relatively short time scales, one
expects that these resonances play a minor role in the long-
term evolution of space debris'. We will show that this is
partially true. Though the regions where SRP resonances are
effective are indeed narrow, for high values of area-to-mass
ratio, they represent the main mechanism to drive a satellite
back to the Earth also when the LEO is sufficiently high to
have a negligible interaction with the residual atmosphere,
e.g., above an altitude h (cid:39) 1000 km. On the other hand, if
we consider low values of area-to-mass ratio, the combined
effect of the atmospheric drag and the SRP resonances can
reduce significantly the lifetime of the object.
One of the fundamental issue to mitigate the space de-
bris problem and control the growth of the debris population
is to see if there exist natural perturbations which facilitate
the reentry to the Earth or, when this is not possible, ensure
a stable graveyard orbit in the long-term. At altitudes when
the atmospheric drag is not effective, the only way to lower
the pericenter altitude is to obtain an eccentricity increase.
For MEO, like those of the GNSS constellations, the grav-
itational lunisolar perturbations can help to this end too
(Rosengren et al. 2015; Alessi et al. 2016). In LEO, apart
from the atmospheric drag, the main natural effects that
can be exploited are the lunisolar gravitational resonance,
in particular the one corresponding to (cid:219)Ω + 2 (cid:219)ω (Alessi et
al. 2017a,b) and the SRP. In particular, as just mentioned,
area-to-mass values representing small spacecraft equipped
with a solar sail can experience a change as dramatic as to
reenter to the Earth, even without the support of the at-
mospheric drag. These situations occur in correspondence
of the (cid:219)Ω + (cid:219)ω − nS resonance, considered as dominant in the
past, but not only. On the other hand, understanding how
the eccentricity might change due to the SRP is crucial also
to control the stability of a graveyard orbit.
As a final note, we acknowledge that looking for the
values of area-to-mass ratio ensuring a reentry, given initial
semi-major axis and inclination, Colombo & de Bras de Fer
(2016) found numerically the same reentry corridors repre-
sented by at least two of the six SRP resonances, that will
be described here.
Table 1. Resonance ψj = n1Ω + n2ω + n3λS in terms of n1, n2, n3.
j
1
2
3
4
5
6
n1
1
1
0
0
1
1
n2
1
-1
1
1
1
-1
n3
-1
-1
-1
1
1
1
2 THE MODEL
Let us assume that the radiation coming from the Sun is
directed normally to the surface of the spacecraft, that is,
the so-called cannonball model. Moreover, let us assume the
orbit of the spacecraft entirely in the sunlight, and the effect
of Earth's albedo as negligible.
Neglecting the light aberration, the solar radiation pres-
sure is a conservative force. The corresponding disturbing
potential, averaged over the orbital motion of the spacecraft,
can be written as (e.g., Krivov et al. 1996)
RSRP =
=
A
m
3
ae[cos ω (cos Ω cos λS + sin Ω sin λS cos )
2 PCR
+ sin ω (cos Ω cos i sin λS cos − sin Ω cos i cos λS)
+ sin i sin ω sin λS sin ],
3
2 PCR
T4 cos ψ4 + T5 cos ψ5 + T6 cos ψ6],
ae[T1 cos ψ1 + T2 cos ψ2 + T3 cos ψ3 + (1)
A
m
where λS is the longitude of the Sun measured on the ecliptic
plane, = 23.439◦ is the obliquity of the ecliptic, P the solar
radiation pressure , CR the reflectivity coefficient, A/m the
area-to-mass ratio,
ψj = n1Ω + n2ω + n3λS,
(2)
with n1 = {0, 1}, n2 = ±1, n3 = ±1, according to j, following
Table 1, and
(3)
.
2
2
T1 = cos2(cid:16)
T2 = cos2(cid:16)
T5 = sin2(cid:16)
T6 = sin2(cid:16)
1
T3 =
2
T4 = −1
2
2
2
(cid:17)
(cid:17)
(cid:17)
(cid:17)
sin() sin(i)
sin() sin(i)
(cid:18) i
(cid:18) i
2
(cid:19)
(cid:19)
2
(cid:18) i
(cid:18) i
2
(cid:19)
(cid:19)
2
cos2
sin2
cos2
sin2
2 both in the eccentricity of the body and in the eccentricity of
the mean apparent solar motion.
If we assume that the dynamics is driven by a single
resonance, then we can write the variation in the orbital
elements as due to only one, say j, of the terms in (1) at a
MNRAS 000, 1 -- 9 (2017)
SRP Resonances in LEO 3
Figure 1. On the top, we show the location of the six main SRP
resonances as a function of i, a for e = 0.01. On the bottom, a
close-up in the LEO region, defined here up to h = 3000 km, in
order to account also for possible graveyard orbits. The curves
were computed assuming (cid:219)Ω = (cid:219)ΩJ2 and (cid:219)ω = (cid:219)ωJ2 . Green: (cid:219)Ω + (cid:219)ω −
nS = 0. Cyan: (cid:219)Ω− (cid:219)ω−nS = 0. Orange: (cid:219)ω−nS = 0. Yellow: (cid:219)ω+nS = 0.
Blue: (cid:219)Ω + (cid:219)ω + nS = 0. Red: (cid:219)Ω − (cid:219)ω + nS = 0.
+ (cid:219)ΩS R P and (cid:219)ω = (cid:219)ωJ2
Figure 2. Location of the six main SRP resonances as a function
of i, a for e = 0.01 (top) and e = 0.1 (bottom) in the LEO region
for prograde orbits for A/m = 1 m2/kg. The curves were computed
assuming (cid:219)Ω = (cid:219)ΩJ2
+ (cid:219)ωS R P . Green: (cid:219)Ω+ (cid:219)ω−nS =
0. Cyan: (cid:219)Ω − (cid:219)ω − nS = 0. Orange: (cid:219)ω − nS = 0. Yellow: (cid:219)ω + nS = 0.
Blue: (cid:219)Ω + (cid:219)ω + nS = 0. Red: (cid:219)Ω − (cid:219)ω + nS = 0. For a given resonance,
the middle curve corresponds to ψ = 90◦. If n3 = 1 the right
curve corresponds to ψ = 0◦, the left one to ψ = 180◦; if n3 = −1
viceversa.
time. By applying the Lagrange planetary equations, we get
Tj
1 − e2
na
√
e
∂cos ψj
∂ω
Tj
1 − e2 sin i
(cid:18) ∂cos ψj
∂Ω
(4)
(cid:19)
− cos i
∂cos ψj
∂ω
√
= −3
A
2 PCR
m
= −3
A
2 PCR
m
na
+ (cid:219)ΩSRP
(cid:219)ΩJ2
(cid:219)ωJ2
+ (cid:219)ωSRP,
=
=
de
dt
di
dt
dΩ
dt
dω
dt
on the spacecraft is the oblateness of the Earth. We have
(cid:219)ΩJ2
(cid:219)ΩSRP =
(cid:219)ωJ2
J2r2n
= −3
a2(1 − e2)2 cos i
2
3
2 PCR
J2r2n
(cid:32)√
3
a2(1 − e2)2
4
3
2 PCR
1 − e2
nae
A
m
A
m
(cid:16)
√
na
=
e
∂Tj
∂i
(cid:17)
1 − e2 sin i
5 cos2 i − 1
(5)
cos ψj
(cid:33)
,
cos ψj
Tj cos ψj −
∂Tj
(cid:219)ωSRP =
∂i
r the equatorial radius of the Earth. The effect of luniso-
where J2 is the second zonal term of the geopotential and
lar perturbations on (cid:219)Ω, (cid:219)ω can be neglected, following, e.g.,
e cos i
√
1 − e2 sin i
na
where n is the mean motion of the satellite and we have
assumed that the only other effect exerting a perturbation
Milani et al. (1987).
Notice that long-period and secular effects associated
MNRAS 000, 1 -- 9 (2017)
4
E. M. Alessi et al.
Figure 3. Resonance width, computed for e = 0.01 and A/m = 1 m2/kg. The middle and right panels show the regions where the width
computed is the largest: they correspond to resonance #1 and resonance #3, respectively.
with the oblateness of the Earth affect only the behaviour
of the longitude of the ascending node, of the argument of
pericenter and of the mean anomaly. That is, the eccentricity
does not change in the long-term due to J2 (see, e.g., Roy
(1982)).
3 THE ECCENTRICITY RESONANCES
From Eq. (4), we can recognise the following 6 resonances
affecting the behaviour in eccentricity
(cid:219)ψj = n1 (cid:219)Ω + n2 (cid:219)ω + n3nS = 0.
(6)
In Fig. 1 we show their location as a function of the incli-
nation and semi-major axis, for given values of e and A/m,
assuming that only the oblateness of the Earth is responsi-
ble of a variation in Ω and ω. This approximation can be
considered valid for A/m = 0.012 m2/kg, which represents
the average value of the intact objects orbiting in LEO. On
the bottom panel, we show a close-up in the LEO region,
defined here up to h = 3000 km, in order to account also
for possible graveyard orbits. Note that we prefer to repre-
sent the resonances as a function of i, a instead of a, e, as
done, e.g., in Colombo & McInnes (2012), because we focus
on values of eccentricity which represent orbits of opera-
tional spacecraft in LEO. Realistic values cannot exceed the
value e = 0.15, the majority of the population in LEO hav-
ing actually e < 0.02. In general, except in the cases where
the spacecraft is orbiting in a region where the atmospheric
drag is dominant, or at a given resonance affecting the ec-
centricity evolution, the eccentricity can be assumed to vary
in a negligible way with respect to the initial value. Also,
in case of resonance, the timescale variation of the eccen-
tricity is much slower than the typical period of the orbits
considered.
In Fig. 2, we show the location of the same resonances
for two different values of eccentricity, accounting also for
the variation of Ω and ω due to SRP. In this case, the loca-
tion of the resonances depends also on Ω, ω, λS. In the Figure
we compute the curves corresponding to ψ = 0◦, 90◦, 180◦ for
A/m = 1 m2/kg which represents a feasible value, given the
current level of technology, for a small spacecraft equipped
with a SRP enhancing device (see Colombo et al. (2017)).
For a given resonance, the middle curve corresponds to the
case ψ = 90◦, which is equivalent to the case of Fig. 1. The
displacement among the curves of a same resonance can be
appreciated at higher LEO, for quasi-circular orbits for the
first and the second resonance, following the convention in
Table 1. Notice, however, that the first equation in (4) states
that ψ = 0◦ and ψ = 180◦ are stationary points for the eccen-
tricity, and thus we expect to have a quasi-secular behaviour
in eccentricity only at approaching the condition ψ = 90◦.
In the same Figures, we note that resonances of different
nature can cross each other. Focusing on prograde LEO,
there exist two main overlapping between SRP resonances.
For quasi circular orbits, i.e., e = 0.01, we have
nS = 0 at a ≈ r + 2193 km and i ≈ 56.06◦3;
• an overlapping between (cid:219)ψ = (cid:219)ω − nS = 0 and (cid:219)ψ = (cid:219)Ω + (cid:219)ω +
nS = 0 at a ≈ r + 1180 km and i ≈ 69◦.
• an overlapping between (cid:219)ψ = (cid:219)ω + nS = 0 and (cid:219)ψ = (cid:219)Ω − (cid:219)ω +
the first crossing occurs at a ≈ r + 2245 km, the second
at a ≈ r + 1220 km. From preliminary simulations, the
Increasing the eccentricity, the value of the semi-major axis
at which the crossing appears also increases, e.g., for e = 0.1
dynamics at the overlapping does not manifest a chaotic
behaviour, rather the two resonances concur to a possible
increase in eccentricity. In general, we can observe isolated
points where the Lyapunov time is significantly lower than
that in the neighbourhood. They do not affect the overall
dynamics, but they can be, however, object of further dedi-
cated studies.
To compute the width of the resonances considered, it
can be applied the same argument developed by Daquin et
al. (2016) for lunisolar doubly-averaged gravitational per-
turbations. Let us assume that the Hamiltonian function of
the system is
+ HSRP,
H = HK + HJ2
where HK is the Keplerian part, HJ2 the one associated with
the oblateness of the Earth, and HSRP = −RSRP. The curves
defining the boundaries of a given resonance j at X∗ ≡ (e∗, i∗)
are defined by the condition
(7)
√
n1 (cid:219)Ω + n2 (cid:219)ω + n3nS = ±2
ν
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:16) ∂2H s e c
3
2 PCR
(cid:17)
∂X2
A
m aeTj
X=X∗
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ,
where
ν =
(cid:18) ∂2H sec
(cid:19)
∂X2
,
X=X∗
(8)
(9)
3 Note that at the same inclination it appears also the well-known
lunisolar gravitational resonance (cid:219)Ω + 2 (cid:219)ω.
MNRAS 000, 1 -- 9 (2017)
SRP Resonances in LEO 5
Figure 4. Maximum variation in eccentricity (color bar) that can be obtained along a given SRP resonance as a function of i, a for
e = 0.01, Ω = 0◦, ω = 0◦, λS = 90.086◦. Left: A/m = 1 m2/kg. Right: A/m = 0.012 m2/kg. See further details in the text.
as in Daquin et al. (2016). For the sake of clar-
and ∂2H s e c
∂X2
ity4,
∂2H sec
∂X2
=
3
2
J2r2
a2(1 − e2)5/2
(cid:16)
(cid:104)
n2
2
(cid:17)
2 − 15 cos2 i
+ 10n1n2 cos i − n2
1
The width computed in this way turns out to be very narrow.
In Figure 3, we provide an example for A/m = 1 m2/kg,
with a close-up in the regions where the width is the largest.
Notice that they correspond to resonances #1 and #3.
With respect to a possible ranking of the six resonances
in terms of the effect they might produce, the maximum
eccentricity variation that can be achieved at a resonance
j can be estimated by integrating the first equation in (4).
This is,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)3
√
1 − e2
na
A
m
Tj(cid:219)ψj
∆e =
2 PCR
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) .
(10)
In Fig. 4 we show the corresponding values as a function of
i, a for e = 0.01, Ω = 0◦, ω = 0◦, λS = 90.086◦ and A/m = 1
m2/kg and A/m = 0.012 m2/kg, sampling the (i, a) plane at
a step of ∆i = 0.2◦ and ∆a = 20 km. In the former case, the
color bar representing the maximum eccentricity variation
achievable is bounded to 0.3, which is the value required to
reenter from the highest LEO considered. In the latter case,
the limit is set to 0.1 just for the sake of clarity, that is,
to be able to appreciate different behaviours, given the ex-
treme narrowness of the resonances. For A/m = 0.012 m2/kg,
it should be noted that the variation that can be obtained
is weak, and that, by definition of resonance, the time to
get to the peak value is significantly long. Also, given that
the range giving the required combination of inclination and
semi-major axis is tapered, a possible exploitation of these
corridors for an operational satellite would require an accu-
rate manoeuvring capacity. A similar argument applies for
resonance #6 for A/m = 1 m2/kg.
Note also that the same expression, Eq. (10), can be
4 Note that n1 are n2 are exchanged in our formulation with re-
spect to the formulation in Daquin et al. (2016).
MNRAS 000, 1 -- 9 (2017)
(cid:105)
.
used to quantify the maximum possible change in eccen-
tricity in the neighborhood of a given resonance, eventually
avoiding an unrealistic operational time. This application is
faced in detail in Schettino et al. (2017). We remark here
that, for the strongest resonances (#1, #2 and #3), the
range of initial inclinations that can be targeted to achieve a
considerable change in eccentricity for the high area-to-mass
case is generally at most ±1◦ with respect to the resonant
value i∗, for given initial semi-major axis and eccentricity.
4 NUMERICAL EVIDENCE
The analysis described in the previous Section was compared
to the results of the numerical simulations performed under
the ReDSHIFT H2020 project (Rossi et al. 2017; Alessi et al.
2017a,b). The aim of the simulations was to map the whole
LEO region in terms of initial orbital elements, considering
two initial epochs and the same two values of area-to-mass
ratio considered here, in order to look for natural deorbit-
ing solutions that may facilitate the compliance with the
international mitigation guidelines. A well-defined grid of
initial conditions was propagated for 120 years by means of
the semi-analytical orbital propagator FOP (Anselmo et al.
1996; Rossi et al. 2009), including in the dynamical model
the geopotential up to degree and order 5, SRP, lunisolar
perturbations and atmospheric drag below an altitude of
1500 km5. The grid in initial inclination, in particular, was
set at steps of 2◦ in the range [2◦ : 120◦].
(a = r + 1200 km, a = r + 1360 km, a = r + 1580
km), for quasi-circular orbits, with a same initial configura-
tion, and assuming A/m = 1 m2/kg. It is possible to appre-
ciate that the eccentricity can experience a non-negligible
In Fig. 5, we show the maximum eccentricity achieved
in 120 years, as a function of initial inclination and eccentric-
ity, starting from three different values of semi-major axis
5 The atmospheric model is based on a Jacchia-Roberts density
model with an exospheric temperature of 1000 K and a variable
solar flux at 2800 MHz (obtained by means of a Fourier analysis
of data corresponding to the interval 1961-1992).
6
E. M. Alessi et al.
variation, only in correspondence of well-defined inclination
bands (the brighter ones). These inclinations are located in
the neighbourhood of the resonant values, described in the
previous Section. In other words, within the limit of the grid
adopted, only at resonant inclination values, the eccentricity
can naturally increase.
It is also interesting to note that in the three examples
the role of the dominant resonance, that is the one providing
the greatest variation, changes. This is what we have called
a resonance switching. In the top panel, this role is taken by
resonance #2; in the middle by resonance #1; in the bottom
by resonances #1 and #3.
In Fig. 6, we compare the maximum eccentricity com-
puted with the numerical simulation with the maximum
variation obtained using estimate (10), for the same three
initial conditions as in Fig. 5, considering e = 0.01 as ini-
tial eccentricity. The analytical value depicted is the sum of
the six variations forecast by Eq. (10) for each resonance.
This choice turned out to approximate better the actual be-
haviour of the maximum eccentricity, especially in the incli-
nation regions between two resonant values. Given that in
the simulations carried out with FOP other perturbations
played a role and that the atmospheric drag was included,
we consider that there exists a reasonable consistency be-
tween the numerical model and the simplified estimation.
The resonance switching can also find an explanation, look-
ing to the relative height corresponding to the cyan points,
in particular for the first two examples.
In general, for all the cases explored, it turned out that
also resonance #4 can yield a significant eccentricity vari-
ation, for the high area-to-mass value. See Fig. 7 for two
examples.
For A/m = 0.012 m2/kg, in Figs. 8 -- 9 we provide two ex-
amples of the eccentricity evolution and the consequent vari-
ation in the pericenter altitude by propagation with FOP, for
altitudes where the atmospheric drag is not effective. They
correspond to quasi-circular orbits, which are located at the
resonance #2 and #1, respectively, at the initial epoch. We
note the variation of hundreds of km obtained, but also the
long time required. In Fig. 10, we show an example when
the SRP resonance aids the action of the atmospheric drag.
Starting from i = 41.95◦, the reentry is achieved in about
55 years; starting from i = 42◦, the lifetime is reduced by
about 8 years. This shows that if the satellite is located in
the neighbourhood of a SRP resonance6, it could be worth to
increase a little the eccentricity to clearing the region more
rapidly. In this way, we can take advantage of both putting
the spacecraft in a denser layer of the atmosphere and the
push in eccentricity given by SRP, which can lower the peri-
center altitude in the long-term. An accurate modelling of
the interaction between SRP and atmospheric drag will be
faced in the future.
5 CONCLUSIONS
In this work, we have reviewed the main orbital resonances
associated with solar radiation pressure which affects the
6 This is, the change in velocity to target the specific inclination
can be afforded.
Figure 5. Maximum eccentricity (color bar) computed over 120
years with FOP, as a function of the initial inclination and eccen-
21, 2020 at 06:43:12 (i.e., λS ≈ 90.086◦). Top: a = r + 1200 km;
tricity for A/m = 1 m2/kg, starting from Ω = 0◦, ω = 0◦ on June
middle: a = r + 1360 km; bottom: a = r + 1580 km.
MNRAS 000, 1 -- 9 (2017)
SRP Resonances in LEO 7
evolution in eccentricity, in order to see if they might be
exploited to design advantageous deorbiting solutions for
Low Earth Orbits. Starting from a well-known theory, we
have showed that at least four out of six resonances can
be considered to increase the eccentricity of a small space-
craft equipped with a solar sail. The variation that can be
achieved along a given resonance, that is, at well-defined val-
ues of semi-major axis, inclination and eccentricity, is such
that a reentry to the Earth can be obtained, even from high
altitudes. For typical values of operational satellites, SRP
resonances can be considered, instead, in combination with
the atmospheric drag, provided an accurate manoeuvring ca-
pability. The simplified model including only solar radiation
pressure and Earth oblateness has been validated against a
more realistic dynamical model. The corresponding results
turn out to be consistent. The novelty of the work regards
the possible applications, but it is also theoretical, because
we have shown the importance of resonances which were ne-
glected in the past.
Future work will look into the role of the initial phas-
ing provided by initial epoch, longitude of ascending node
and argument of pericenter, which could ensure either an
increase or an eccentricity decrease. A detailed analysis on
the characteristic frequencies for the eccentricity evolution in
the resonant regions due to SRP is ongoing. That study will
include also a comparison between the theoretical derivation
explained here and the numerical results, without the con-
tribution of the drag. Part of the results can be found in
Schettino et al. (2017). Also, the inclination evolution was
neglected in the analysis presented, and it will be evaluated
in detail. From the point of view of a space operator, the
model is still simplified in the sense that the orientation of
a sail cannot be assumed as always directed normal to the
solar radiation. This is another issue, which will deserve fur-
ther research.
ACKNOWLEDGEMENTS
This work is funded through the European Commission
Horizon 2020, Framework Programme for Research and In-
novation (2014-2020), under the ReDSHIFT project (grant
agreement n◦ 687500).
We are grateful to Bruno Carli, Camilla Colombo and
Kleomenis Tsiganis for the useful discussions we had. We
thank the anonymous reviewer for the comments that im-
proved considerable the paper. E.M.A. is also grateful to
Florent Deleflie for the period spent in Lille, and to Jerome
Daquin for the suggestions given.
REFERENCES
Alessi E. M., Deleflie F., Rosengren A. J., Rossi A., Valsecchi
G. B., Daquin, J., Merz, K., 2016, Celestial Mechanics and
Dynamical Astronomy, 125, 71
Alessi E. M., Schettino G., Rossi A., Valsecchi G. B., 2017a, LEO
Mapping for Passive Dynamical Disposal, Proceedings of the
7th European Conferences on Space Debris, Darmstadt, Ger-
many
Alessi E. M., Schettino G., Rossi A., Valsecchi G. B., 2017b, Dy-
namical Mapping of the LEO Region for Passive Disposal De-
Figure 6. Maximum eccentricity computed over 120 years with
FOP in purple, and following Eq. (10) in cyan, as a function of
the initial inclination, for A/m = 1 m2/kg, starting from e = 0.01,
Top: a = r + 1200 km; middle: a = r + 1360 km; bottom:
Ω = 0◦, ω = 0◦ on June 21, 2020 at 06:43:12 (i.e., λS ≈ 90.086◦).
a = r + 1580 km.
MNRAS 000, 1 -- 9 (2017)
8
E. M. Alessi et al.
for A/m = 1 m2/kg, starting from a = r + 1460 km (left) and a = r + 1520 km (right). In both cases, Ω = 0◦, ω = 180◦ on June 21, 2020
Figure 7. Maximum eccentricity (color bar) computed over 120 years with FOP, as a function of the initial inclination and eccentricity
at 06:43:12 (i.e., λS ≈ 90.086◦).
Figure 8. On the left, eccentricity evolution computed with FOP for A/m = 0.012 m2/kg, starting from a = r + 1940 km, e = 0.01,
i = 80◦, Ω = 90◦, ω = 90◦ on June 21, 2020 at 06:43:12 (i.e., λS ≈ 90.086◦). On the right, the consequent variation in pericenter altitude.
Figure 9. On the left, eccentricity evolution computed with FOP for A/m = 0.012 m2/kg, starting from a = r + 2600 km, e = 0.01,
i = 36◦, Ω = 90◦, ω = 90◦ on June 21, 2020 at 06:43:12 (i.e., λS ≈ 90.086◦). On the right, the consequent variation in pericenter altitude.
MNRAS 000, 1 -- 9 (2017)
SRP Resonances in LEO 9
bations and satellite geodesy, Adam Hilger Ltd., Bristol and
Boston
Musen P., 1960, Journal of Geophysical Research, 65, 1391
Rosengren A. J., Scheeres D. J., 2013, Advances in Space Re-
search, 52, 1545
Rosengren A., Alessi E. M., Rossi A., Valsecchi G. B., 2015,
Monthly Notices of the Royal Astronomical Society, 449, 3522
Rossi A., Anselmo L., Pardini, C. Jehn, R., Valsecchi, G. B.,
2009, The new space debris mitigation (SDM 4.0) long term
evolution code, Proceedings of the Fifth European Conference
on Space Debris, Darmstadt, Germany, Paper ESA SP-672
Rossi A., and the ReDSHIFT team, 2017, The H2020 Project
ReDSHIFT: Overview, First Results and Perspectives, Pro-
ceedings of the 7th European Conference on Space Debris,
Darmstadt, Germany
Roy A. E., 1982, Orbital Motion, Adam Hilger Ltd., Bristol
Schettino G., Alessi E. M., Rossi A., Valsecchi G. B., 2017, Inter-
national Astronautical Congress, paper IAC-17.C1.9.2.
Touma J., Wisdom J., 1998, The Astronomical Journal, 115, 1653
Valk S., Lemaıtre A., Anselmo L., 2007, Advances in Space Re-
search, 41, 1077
This paper has been typeset from a TEX/LATEX file prepared by
the author.
for A/m = 0.012 m2/kg, starting from a = r + 840 km, e = 0.05,
Figure 10. Pericenter altitude evolution computed with FOP
i = 42◦ (purple) and i = 41.95◦ (green), Ω = 270◦, ω = 270◦ on
June 21, 2020 at 06:43:12 (i.e., λS ≈ 90.086◦).
sign, 2017, International Astronautical Congress IAC-2017,
paper IAC-17.A6.2.7
Anselmo L., Cordelli A., Farinella P., Pardini C., Rossi A.,
1996, Study on long term evolution of Earth orbiting debris,
ESA/ESOC contract No. 10034/92/D/IM(SC)
Beutler G., 2005, Methods of Celestial Mechanics II: Application
to Planetary System, Geodynamics and Satellite Geodesy,
Springer-Verlag, Berlin Heidelberg
Breiter S., 1999, Celestial Mechanics and Dynamical Astronomy,
74, 253
Brouwer D., Clemence G M., 1961, Methods of Celestial Mechan-
ics, Academic Press, New York
Casanova D., Petit A., Lemaıtre A., 2015, Celestial Mechanics
and Dynamical Astronomy, 123, 223
Celletti A., Efthymiopoulos C., Gachet F., Gale¸s C., 2017, Inter-
national Journal of Non-Linear Mechanics, 90, 147
Colombo C., Lucking C., McInnes C. R., 2012, Acta Astronautica,
81, 137
Colombo C., McInnes C. R., 2012, International Astronautical
Congress IAC-2012, paper IAC-12-C1.4.12
Colombo C., de Bras de Fer T., 2016, International Astronautical
Congress IAC-2016, paper IAC-16-A6.4.4
Colombo C., Rossi A., Dalla Vedova F., 2017, International As-
tronautical Congress IAC-2017, paper IAC-17.A6.2.8
Cook G. E., 1962, The Geophysical Journal of the Royal Astro-
nomical Society, 6, 271
Daquin J., Rosengren A. J., Alessi E. M., F. Deleflie, G. B. Valsec-
chi, Rossi A., 2016, Celestial Mechanics and Dynamical As-
tronomy, 124, 335
Frouard J., Fouchard M., Vienne A., 2010, A&A, 515, A54
Hughes S., 1977, Planetary and Space Science, 25, 809
Hughes S., 1980, Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences, 372, 243
Hughes S., 1981, Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences, 375, 379
Krivov A. V., Sokolov L. L., Dikarev V. V., 1996, Celestial Me-
chanics and Dynamical Astronomy, 63, 313
Lantukh D., Russell R. P., Broschart S., 2015, Celestial Mechanics
and Dynamical Astronomy, 121, 171
Lucking C., Colombo C., McInnes C. R., 2012, Acta Astronautica,
77, 197
Milani A., Nobili A., Farinella P., 1987, Non gravitational pertur-
MNRAS 000, 1 -- 9 (2017)
|
1003.0457 | 1 | 1003 | 2010-03-01T21:15:49 | Lack of Transit Timing Variations of OGLE-TR-111b: A re-analysis with six new epochs | [
"astro-ph.EP"
] | We present six new transits of the exoplanet OGLE-TR-111b observed with the Magellan Telescopes in Chile between April 2008 and March 2009. We combine these new transits with five previously published transit epochs for this planet between 2005 and 2006 to extend the analysis of transit timing variations reported for this system. We derive a new planetary radius value of 1.019 +/- 0.026 R_J, which is intermediate to the previously reported radii of 1.067 +/- 0.054 R_J (Winn et al. 2007) and 0.922 +/- 0.057 R_J (Diaz et al. 2008). We also examine the transit timing variation and duration change claims of Diaz et al. (2008). Our analysis of all eleven transit epochs does not reveal any points with deviations larger than 2 sigma, and most points are well within 1 sigma. Although the transit duration nominally decreases over the four year span of the data, systematic errors in the photometry can account for this result. Therefore, there is no compelling evidence for either a timing or a duration variation in this system. Numerical integrations place an upper limit of about 1 M_E on the mass of a potential second planet in a 2:1 mean-motion resonance with OGLE-TR-111b. | astro-ph.EP | astro-ph |
Lack of Transit Timing Variations of OGLE-TR-111b: A
re-analysis with six new epochs1
E. R. Adams2,M. L´opez-Morales3, J. L. Elliot2,4, S. Seager2,4, D. J. Osip5
ABSTRACT
We present six new transits of the exoplanet OGLE-TR-111b observed with
the Magellan Telescopes in Chile between April 2008 and March 2009. We com-
bine these new transits with five previously published transit epochs for this
planet between 2005 and 2006 to extend the analysis of transit timing vari-
ations reported for this system. We derive a new planetary radius value of
1.019 ± 0.026 RJ , which is intermediate to the previously reported radii of
1.067 ± 0.054 RJ (Winn et al. 2007) and 0.922 ± 0.057 RJ (D´ıaz et al. 2008).
We also examine the transit timing variation and duration change claims of
D´ıaz et al. (2008). Our analysis of all eleven transit epochs does not reveal any
points with deviations larger than 2σ, and most points are well within 1σ. Al-
though the transit duration nominally decreases over the four year span of the
data, systematic errors in the photometry can account for this result. Therefore,
there is no compelling evidence for either a timing or a duration variation in this
system. Numerical integrations place an upper limit of about 1 M⊕ on the mass of
a potential second planet in a 2:1 mean-motion resonance with OGLE-TR-111b.
Subject headings: stars: planetary systems -- OGLE-TR-111
1This paper includes data gathered with the 6.5 meter Magellan Telescopes located at Las Campanas
Observatory, Chile.
2Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77
Massachusetts Ave., Cambridge, MA, 02139
3Hubble Fellow; Carnegie Institution of Washington, Department of Terrestrial Magnetism, 5241 Broad
Branch Road NW, Washington, DC 20015-1305
4Department of Physics, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge, MA,
02139
5Las Campanas Observatory, Carnegie Observatories, Casilla 601, La Serena, Chile
-- 2 --
1.
Introduction
Transiting exoplanets provide a wealth of information for studies of the physical param-
eters of planets and their environments. For example, the combination of several accurately
timed transits of a known transiting exoplanet can be used not only to improve estimates
of the planetary radius and orbital parameters of the star-planet system, but also to detect
additional objects. Detecting potential variations of parameters such as the inclination and
duration of the transits would indicate a precesing planetary orbit, potentially caused by
another planet (Miralda-Escud´e 2002). We can also use transit timing to search for addi-
tional planets or moons, as discussed in several recent theory papers (Holman and Murray
2005; Agol et al. 2005; Heyl and Gladman 2007; Ford and Holman 2007; Simon et al. 2007;
Kipping 2009; Kipping et al. 2009). The idea is that the presence of additional objects will
perturb the orbit of the transiting planet, producing transit timing variations (TTVs) or
transit duration variations (TDVs). Those TTVs and TDVs can be detected by monitoring
transits over many orbital periods. The absence of such variations can be also used to place
limits on the mass and orbital parameters of additional objects in those planetary systems
and to gain insight into the systems' architectures.
Recent observations show hints of timing variations for some transiting planets, but no
definitive detection of additional planets or satellites has been reported using this technique.
The most interesting results so far are (1) the absence of TTVs in several systems, which do
not host planets more massive than several Earth masses in low-order resonant orbits (see a
summary of constraints that can be placed in Table 1); (2) the tentative detection of TDVs
in GJ436, roughly 3 minutes per year (Coughlin et al. 2008), a trend consistent with the
presence of a low-mass companion (< 12 M⊕) in a close exterior but non-resonant orbit; this
result is consistent with the 8 M⊕ limit placed by transit timing (Bean and Seifahrt 2008);
and (3) the preliminary detection of TTVs with a maximum residual of 156 ± 48 sec (3.3σ)
over a period of 2 years reported by D´ıaz et al. (2008) for OGLE-TR-111b, the subject of
this paper.
OGLE-TR-111b is a 0.5MJ hot Jupiter orbiting its host star, a faint (I = 15.5) K
dwarf, every 4.01 days. This object was first announced as a transiting planet candidate
by Udalski et al. (2002), and was confirmed to have planetary mass by Pont et al. (2004).
The physical parameters of the planet were refined over the next two years, with several
new radial velocity measurements (Gallardo et al. 2005; Silva and Cruz 2006; Santos et al.
2006). The first high precision transit photometry was provided by Winn et al. (2007), with
two I -band transits of the planet on 2006 Feb 21 and Mar 5. Shortly after, Minniti et al.
(2007) published a V -band transit from 2005 April 9 and noted that the midtime occurred
5 minutes earlier than expected from the ephemeris in Winn et al. (2007), although with
-- 3 --
only three epochs they could draw no firm conclusions. A follow-up paper by D´ıaz et al.
(2008) reported two consecutive I -band transits of OGLE-TR-111b on 2006 Dec 19 and 23.
Combining all five epochs, they concluded that the previously claimed TTVs were real, with
the residuals spanning −156 ± 48 to +98 ± 39 seconds. Among other scenarios, they noted
that if OGLE-TR-111b were in an eccentric orbit with e ∼ 0.3, the observed TTVs would
be consistent with the presence of an Earth-mass planet near an exterior 4:1 resonant orbit.
Additionally, D´ıaz et al. (2008) noted two parameters with marginally discrepant values
across the five transits (see Table 2). Compared to the results from Winn et al. (2007), the
D´ıaz et al. (2008) values for the planetary radius disagreed at the 10% level, or 1.3σ, and
the total transit duration differed by 1.6σ. The radius ratio discrepancy was suggested to
be the result of the parameters chosen for the image subtraction photometry, which focused
on precise timing rather than on an accurate transit depth determination. The duration
variation, if real, could be due to a perturber decreasing the orbital inclination, which would
offer another way of determining the properties of the third body in the system suggested
by their TTVs.
Here we present six new transits observed during 2008 and 2009, which double the
number of high-quality transit light curves available for OGLE-TR-111b. In § 2 we describe
the collection and analysis of the new data. In § 3 we describe the transit model fitting, and
discuss additional sources of error not included in the formal fit. In § 4 we combine the six
new transits with the five previously published observations and provide a new analysis of
parameter variation in the OGLE-TR-111 system. In § 5 we discuss the implications of our
results.
2. Observations and Data Analysis
All six new transits were observed between April 2008 and March 2009 in the Sloan i′
filter with the new MagIC-e2v camera1 on Magellan. The MagIC-e2v camera has a field of
view of 38′′ x 38′′ and a plate scale of 0.037′′ per pixel unbinned. With such high resolution
and good average seeing at the site, blends are minimized and aperture photometry can
be successfully applied even in fairly crowded fields. The camera can be operated in two
different modes: single exposure mode, with a readout time of about 5 seconds per exposure,
and frame transfer mode, with a readout time of only 3 milliseconds between frames in an
1The MagIC-e2v detector, which shares a dewar with the older SiTe CCD, is identical to the red CCD
on HIPO, a fast read-out direct imaging camera and one of the first generation instruments to be flown
on SOFIA; both cameras use the LOIS control software (Dunham et al. 2004; Taylor et al. 2004; Osip et al.
2008).
-- 4 --
image cube. Our first four transits were observed in single exposure mode. The frame
transfer mode first became available after engineering in July 2008, and was used for the
last two transits of OGLE-TR-111b. The gain and read noise of the first four transits were
2.4 e-/ADU and 5.5 e- per pixel, respectively; after engineering, these values were changed
to the current values of 0.54 e-/ADU and 5 e- per pixel.
The exposure times during each transit were adjusted to maintain a minimum count
level of about 106 integrated photons, both for the target and multiple nearby comparison
stars. For the 2008 transits, we collected unbinned (1x1) data with exposure times between
30 and 120 sec, depending on the observing conditions, with an additional readout overhead
per exposure of 5 seconds per frame. The 2009 data were collected in frame-transfer mode
with the camera binned 2x2, which yielded an improved sampling rate of 15-30 sec per frame.
Details of the observing settings are noted in Table 3.
Accurate timing is of the utmost importance for this project, so special care was taken to
ensure that the correct times were recorded in the image headers. For the 2008 transits, the
start times for each image were recorded from a network time server, which was verified by
eye to be synchronized with the observatory's GPS clocks at the beginning of each night. For
the 2009 observations, the times came from a PC104 (a small embedded control computer),
which received unlabeled GPS pulses every second. As with the network time server, the
PC104 was synchronized with the observatory's GPS before each transit observation.
In
both cases the time signals written to the image headers agree within one second with the
GPS time. During both 2009 transits, a software failure caused the times for a few image
cubes to not be recorded in the headers of the images, but we were able to reconstruct
the observation times with precisions better than a second from detailed system logs. One
second is a conservative estimate of the intrinsic error for the start time for each frame, and
is significantly smaller than the mid-transit times errors.
2.1. Data Analysis
All data were overscan corrected and flattened using IRAF.2 The photometry was per-
formed using the IRAF routine phot, part of the apphot package. Depending on the binning
applied and the seeing during each transit, a subset of apertures between 6-25 pixels in ra-
dius were examined for the target star and each comparison star. Between one and seven
2IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the As-
sociation of Universities for Research in Astronomy, Inc., under cooperative agreement with the National
Science Foundation.
-- 5 --
stars were selected as comparisons from the list of 10-20 field stars present in each frame.
The comparison stars had to pass several selection criteria: to be similar in brightness to
the target, to not be blended, and to not be variable. The best apertures were identified as
the ones which yielded the smallest scatter in the out-of-transit flux. The choice of position
of the sky background annulus also influenced which stars could be used for comparison; see
§ 3.3 for a discussion of systematic errors resulting from aperture settings. We explored sky
regions with inner radii from 20-40 pixels and 10-30 pixels width, and selected the one which
(a) provided several comparison stars of suitable brightness that were not variable, and (b)
produced the lowest noise in the out-of-transit baseline while not introducing spikes or other
obvious problems in the transit light curve.
In each of the resultant light curves, the out-of-transit baseline was examined for linear
correlations with several variables: airmass, seeing, telescope azimuth, (x,y) pixel location,
and time since beginning of transit. The parameters chosen are either directly correlated
with physical phenomena that affect photon rates (e.g.
seeing, airmass), or are proxies
for other effects (e.g. the telescope azimuth is correlated with the de-rotator rates, which
were not recorded for several transits). Only the 2009 transits exhibited significant trends,
against telescope azimuth (both) and seeing (20090313). For each transit, we corrected
those trends by successively subtracting linear fits to each variable. Detrending was critical
in producing usable light curves for these two transits, but may also have introduced smaller-
order systematic effects, particularly in the transit depth (see also § 3.3).
We now briefly describe the observations and the photometric reduction of each transit
dataset. All transit fluxes and times are available online; an excerpt is shown in Table 4.
2.1.1. 20080418
Transit 20080418 was observed during engineering time just after the e2v CCD was first
installed on the telescope; due to engineering constraints, only the second half of the transit
was observed. The field was repositioned before egress to eliminate diffraction spikes from
a nearby bright star by moving it further off-chip. The airmass was low and fairly constant
(1.2-1.3) and the seeing was good, increasing slightly from 0.4′′ to 0.5′′ during the transit.
The out-of-transit data showed no apparent trends with the parameters checked.
-- 6 --
2.1.2. 20080422
Transit 20080422 was observed on an intermittently cloudy night with highly variable
transparency, with counts on the target star varying by a factor of 6 within a few frames.
We found that eliminating the lowest count frames, those with fewer than 300,000 counts
on the target star, significantly decreased the scatter of the light curve. The seeing ranged
from 0.5-0.6′′, and the airmass was low and fairly constant (1.2-1.3). The out-of-transit data
showed no apparent trends with the parameters checked.
2.1.3. 20080512
Transit 20080512 had stable photometric conditions for the entire pre-transit baseline.
During the transit there were two drops in target counts (by a factor of 2) that coincided
with sudden seeing jumps (0.4′′ spiking to 0.6′′). The field also drifted substantially (by
∼100 rows and ∼100 columns) due to tracking problems; about 30 minutes of post-transit
baseline had to be discarded because of strong image elongation. The airmass ranged from
1.2 to 1.7. The out-of-transit data showed no apparent trends with the parameters checked.
2.1.4. 20080516
Transit 20080516 had very stable photometric conditions for most of the transit. The
seeing gradually increased from 0.4′′ to 0.6′′, and the airmass ranged from 1.2 to 2.0. The star
also drifted substantially toward the end of transit (by ∼ 200 rows and ∼ 200 columns) for
unknown reasons. The out-of-transit data showed no apparent trends with the parameters
checked.
2.1.5. 20090217
Transit 20090217 was the first of OGLE-TR-111b to be observed with the new frame
transfer mode. The seeing fluctuated from 0.7-1.1′′, while the airmass decreased from 1.8 to
1.2. This transit was detrended for a slope correlated with the telescope azimuth.
-- 7 --
2.1.6. 20090313
Observations for transit 20090313 began late due to telescope problems and thus there is
no pre-transit baseline. Seeing conditions were initially very poor, spiking to 2′′, but improved
substantially during the second half of the transit, to around 0.6′′. The airmass ranged from
1.2 to 1.5. The huge disparity in seeing made photometry on this transit challenging, and
the best light curve resulted from using a small aperture around the target and only the
brightest single comparison star, which saturated for a few frames that had to be discarded.
This transit was detrended for slopes correlated with the seeing and the telescope azimuth.
2.1.7. Literature light curves
To eliminate any uncertainty in comparing our transits to previously published transits,
which might result from different models, fitting procedures, or fixed parameter values, we
have obtained tables of the times and fluxes for each of the five transits drawn from the
literature: 20050409 (Minniti et al. 2007), 20060221 and 20060305 (Winn et al. 2007), and
20061219 and 20061223 (D´ıaz et al. 2008). We have used the original photometry, except for
converting from magnitudes into fluxes (where applicable) and converting the mid-exposure
times from Heliocentric Julian Day, or HJD, into Barycentric Julian Day, or BJD, to be
consistent with the rest of our analysis; the difference between the two time systems is much
smaller than our errors, though, a few seconds at most. All values presented for the literature
light curves are taken from our re-analysis of the published photometry using our model and
fitting, unless otherwise noted.
The light curves for all six new transits and the five literature transits are shown in
Figure 1, together with the best joint model fit, which will be described in detail in § 3.
3. Transit fitting results
3.1. Model
Each light curve was fitted with the Mandel and Agol (2002) algorithm to generate an-
alytical models, using the basic optimized model-fitting code described in Carter and Winn
(2009), but without the wavelet analysis. In the models we assumed that OGLE-TR-111b
has zero obliquity, oblateness and orbital eccentricity. We used a quadratic limb darkening
law of the form
√1 − r2) − u2(1 −
√1 − r2)2,
I(r) = 1 − u1(1 −
(1)
-- 8 --
with the initial parameters for u1 and u2 set to the values for the appropriate filter (Claret
2000, 2004). Although analyses of different limb darkening laws have shown that using a
non-linear limb darkening law is important (Southworth 2008), it is generally not possible to
fit both quadratic limb-darkening coefficients except on the highest quality, typically space-
based, data. We thus fixed the quadratic term u2 and only fit for the linear term, u1. (We also
fixed u1 for the sparsely-sampled transit on 20050409, the only one observed in V band.) The
values for u1 and u2 are calculated using the jktld program by Southworth (2008)3, assuming
T = 5044 K, log g = 4.25 cm/s2, [M/H] = 0, and Vmicro = 2 km/s. We used the limb
darkening values corresponding to the Sloan i′ filter for the new data, the VC filter for the
transit from Minniti et al. (2007), and the IC filter for the transits from Winn et al. (2007)
and D´ıaz et al. (2008), as listed in Table 5 (though we note for completeness that the actual
filters used in the literature light curves, Mould V, CTIO I and Bessel I, respectively, do not
correspond precisely to the Cousins V and I filter parameters that were available). We fixed
the orbital period to P = 4.01445 days; later experiments with slightly different values had
little effect. The other free parameters in the model are the radius ratio, k, inclination, i,
semimajor axis in stellar radii, a/R∗, out-of-transit flux, FOOT and transit midtime, TC. We
assume throughout the fits that M∗ = 0.81 M⊙, R∗ = 0.83 R⊙, and Mp = 0.52 MJ , based
on the spectroscopic work of Santos et al. (2006).
3.2. Light curve fits
To determine the best fit value and error of each model parameter, we used a Monte
Carlo Markov Chain (MCMC) method, as described in Carter and Winn (2009). The initial
values for each parameter were computed by a joint least squares fit to each light curve
independently. We then weighted each light curve by the reduced χ2 of this fit so that the
new reduced χ2 = 1. (By doing this we are assuming the transit model is correct in order to
determine the error on each transit, rather than assuming a noise model, e.g. photon noise,
in order to test the transit model.) A joint least squares fit of all weighted transits had a
reduced χ2 = 1.05. Starting from the initial least squares values, we constructed chains of
1,000,000 links, where the acceptance rate for each parameter is between 20-60%. We fit
all eleven transits simultaneously, assuming common values for k, i, a/R∗, and u1,x (where
x is the appropriate filter), with u2,x fixed, but fitting each transit for its own FOOT and
TC. We did not fit for an airmass slope (see § 3.3. The first 50,000 points of the MCMC
were discarded to eliminate bias toward the initial conditions. We created three independent
MCMC chains, checked that the Gelman-Rubin statistic (Gelman and Rubin 1992) is close
3http://www.astro.keele.ac.uk/ jkt/codes/jktld.html
-- 9 --
to 1 to ensure convergence, and then combined the chains to determine the distribution of
all parameters, including the total duration of the transit, T14 (the time from first to fourth,
or final, contact), and the impact parameter, b, which are derived from fitted parameters.
We plot the best model fit with the data in Figure 1 and tabulate the fit results in Table 5,
where we report for each parameter the median value and the 68.3% credible interval (the
equivalent to a 1σ standard deviation if the distribution is Gaussian). The distributions for
each parameter are shown in Figure 2.
The new radius ratio for OGLE-TR-111b based on an analysis of all eleven light curves
yields a planetary radius Rp = 1.019 ± 0.007 RJ , if we consider only the formal fit errors;
accounting for the error on the stellar radius, which is now the dominant source of error, we
find a more realistic error bar is Rp = 1.019± 0.026 RJ . Note that if we only use the six new
light curves, which have more consistent radii, the value for the radius ratio is very similar
(formal fit Rp = 1.015 ± 0.009 RJ , or Rp = 1.015 ± 0.026 RJ with stellar errors).
As a test of the robustness of our parameter determination, we also ran additional
MCMC fits for each transit independently, with results in Table 6. Although most of the
parameters agree within the formal 1σ errors between the individual and joint fits, there are
some notable exceptions. In § 4.2 we investigate variability (previously noted by D´ıaz et al.
(2008)) in both the radius ratio, k, and the total transit duration, T14.
3.3. Systematic Errors and Correlated Noise
One of the most apparent results from the fits to individual light curves was that the
radius ratios are similar for transits observed on the same instrument and reduced by the
same group. This may indicate a degree of subjectivity in the light curve generation process,
both from the choice of photometry method (e.g. aperture, image subtraction, deconvolution,
etc.) and from the specific choice of reduction parameters (e.g. aperture size and sky region
for aperture photometry). These choices can result in systematic errors in the transit depth,
particularly when comparing transits from multiple sources. Both image subtraction and
aperture photometry require fine-tuning a number of parameters, and there is no single
prescription for how to get the absolute best light curve: the same method applied to the same
transit could produce similar quality light curves, as measured by the scatter of residuals or
out-of-transit flux, which nonetheless differ in depth by more than the formal fitted errors.
It has been noted by Winn et al. (2007) that with image subtraction, slight changes in
both the difference flux and the reference flux can cause the measured radius ratio to vary
by a few percent of its value, although their estimate of that effect on their own data,
δk = 0.0002, is much less than our formal fit error of 0.0008. This effect was also alluded
-- 10 --
to by D´ıaz et al. (2008) as an explanation for their shallow depths compared to previous
results, although they did not provide a numerical estimate of the magnitude of this effect.
An analysis by Gillon et al. (2007) of a different transiting planet, OGLE-TR-132b, found
that image subtraction is particularly prone to misestimating the transit depth, compared to
the alternative methods of aperture and deconvolution photometry. This effect for OGLE-
TR-132b in their data causes the radius ratio to differ by 1-2% depending on the choice of
parameters. If similar levels of error were present for OGLE-TR-111b, particularly in the
D´ıaz et al. (2008) curves which were acknowledged to not be optimized toward finding the
correct depths, the systematic error on the radius ratio would be 0.0013−0.0025, comparable
to the formal fit error of 0.002 on the individual curves. Assuming a median value of 0.002
for the systematic error and adding it in quadrature with the formal fit error, a better error
estimate on the radius ratio would be 0.003.
We attempted to quantify the systematic error for aperture light curve generation as
follows. For each of our light curves, we used 4-5 sets of apertures, sky radii and widths,
and different comparison stars, with a goal toward minimizing the scatter in the out-of-
transit baseline. The choice of comparison stars in particular depends on the aperture and
sky choices, because certain stars are usable under some choices but not others, especially
stars of very different brightnesses. Additionally, we examined the effect of detrending light
curves, which must be done carefully because systematic trends in the data can distort the
measured radius, but so can an incorrectly-removed slope. We only detrended transits which
had strong slopes in the out of transit baseline or otherwise had distorted shapes, and then
only for the parameters which best corrected the shape defects. It is possible that residual
trends against parameters we did not consider, or slightly nonlinear trends, could remain
in the data and distort our estimate of the radius; we chose to stick with linear trends
against a few meaningful physical parameters to avoid introducing unnecessary complexity.
To test whether any important slopes remained, we ran a joint MCMC fit which includes a
differential extinction term (a trend with airmass), but found that the fitted airmass slopes
were slight and the difference in resulting parameters were in all cases less than the formal
1σ errors. The results of our explorations of both the aperture and sky choices and the
detrending parameters are that we can produce light curves with similar shapes and scatter,
and that the radius ratios vary by 0.001-0.004 for an individual transit. (Some transits are
much more resilient to parameter choices than others.) Thus, for aperture photometry also
the systematic error on the radius ratio based solely on the parameter choice is of order the
formal fit errors.
To estimate the amount of correlated noise in the light curves, we used two methods.
The first is residual permutation, which shifts the residuals for each transit through every
point in time and adds it to the best model fit; we also assumed time invariance and reversed
-- 11 --
the residuals, then permuted again, for a total of 214-1600 curves for each transit. We fit
a least-squares transit model to each permuted curve for all eleven transits. For the radius
ratio and the transit midtime, the errors from the residual permutation method for both
values were greater by a factor of 1-3, depending on the light curve (note that this factor is
not the same for each parameter; see Table 6). We got similar results when we ran a joint
fit of all 11 transits with 10,000 curve ensembles, randomly selecting for each transit one of
its individually permuted light curves.
An alternative way to estimate the error contributed by correlated noise is to calculate
how the noise scales with time averaging (Pont et al. 2006). We calculate the standard
deviation on the residuals in bin sizes from 10-30 minutes and compare that value to what
we would expect if the noise behaved like Poisson noise (i.e., a decrease in the noise with
√N points). We calculate the amount by which the real noise is greater than the estimated
noise, and find that it is greater by a factor of 1.5-3 times the purely Poisson noise level,
depending on the transit. The increased noise factors agree with the values found by residual
permutation, and for simplicity we use the scaled errors from the residual permutation
throughout.
For all of these reasons, the formal fit errors reported in Table 5 are underestimated due
to both systematic errors and correlated noise. The errors we adopted have been inflated
based on the residual permutation method. This means that, once systematics are included,
neither the observed timing variations for 20050409 and 20080418 nor the duration variation
are statistically significant (see § 4.1 and § 4.2).
4. Results
4.1. Timing
The central midtimes for all 11 transits that we fit are summarized in Table 5 and
illustrated in Figure 4. Recently Pietrukowicz et al. (2010) have reanalyzed the photome-
try for 30 transits of OGLE planets and planet candidates, among them OGLE-TR-111b,
and they have found a different midtime than originally reported in Minniti et al. (2007):
TC,new = 2453470.5676 ± 0.0005, compared to the original published value of TC,orig =
2453470.56413 ± 0.00067, a difference of 300 seconds. (It is not clear what is the source
of such a large shift, but one possibility is a mistake in the UTC-BJD correction.) Signifi-
cantly, the new time is much closer to the expected time of transit.
Another potential pitfall when comparing times from multiple groups has recently noted
by Eastman and Agol (in prep). Most researchers, and indeed most common conversion tools
-- 12 --
(e.g. barycen.pro in IDL and setjd in IRAF) by default omit the correction from UTC to
TT, which in 2009 was 66.184 seconds. We have confirmed that the times published by
Pietrukowicz et al. (2010) and Winn et al. (2007) do not account for the UTC-TT correc-
tion (personal communications), and we assume that the times in Minniti et al. (2007) and
D´ıaz et al. (2008) likely did not either. We have therefore added the appropriate correction
to the reported BJD times for these light curves. (Note that the smaller order deviations
introduced by using UTC rather than TT times in calculating the BJD correction terms are
at most a few seconds, and for this work those deviations fall well within the timing errors;
however, with higher precision data on other systems it would be very important to consis-
tently calculate the BJD times.) All of the transit midtimes in Table 5 have been corrected
to the BJD-TT system; additionally, we have added 64.184 seconds to the Pietrukowicz et al.
(2010) midtime to get TC = 245470.56834 (BJD), the value used in all subsequent calcula-
tion.
The top panel of Figure 4 shows the ephemeris from D´ıaz et al. (2008), derived from
the first five transits, along with the timing residuals for all eleven transit based on our
joint-fit values (Table 5). To correct for the linear drift in the residuals, we calculate a new
constant-period transit ephemeris, omitting the half-transit (20080418), which has very large
errors, and using the new Pietrukowicz et al. (2010) time for 20050409, and find:
TC = 2454092.80717(16)[BJD] + 4.0144463(10)N.
(2)
where TC is the predicted central time of a transit, and N is the number of periods since
the reference midtime, and the values in parentheses are errors on the last digits. We find
almost identical ephemeris values if we use all transits and if we use the original time for
20050409, although the errors are several times greater in both cases. Our adopted fit has a
reduced χ2 = 0.5.
The lower panel in Figure 4 shows the new ephemeris and the timing residuals. (Note
that the same times are used in both panels, and only the ephemeris has changed.) With 1σ
errors ranging from 36 s to 114 s, only the original 20050409 time is more than 2σ from zero,
and of the other transits only the half transit 20080418, which is inherently less trustworthy,
is more than 1σ. Thus, we conclude that the timing deviations reported by D´ıaz et al.
(2008), which depended heavily on the old time for 20050409, do not exist, and we see no
evidence for timing variations in our data.
-- 13 --
4.2. Parameter variation
D´ıaz et al. (2008) found that their value for the total duration was 4.4 minutes shorter
than that found by Winn et al. (2007), a 1.6σ result given the respective quoted errors. If
this decrease is real, it would be of great interest, since a likely explanation would be that the
inclination of OGLE-TR-111b is precessing, possibly due to the presence of another planet.
On the other hand, the variation could be due to errors in the photometry or undetected
correlated noise. Our values for the best fit duration for each transit are plotted in Figure 3
and tabulated in Table 6, along with the parameters the duration was derived from: i, k,
and a/R∗. Note that the errors in this table have been increased from the formal fit errors
by a factor derived from the residual permutation method discussed in § 3.3.
At first glance there does appear to be a decrease in duration over time. To compare
the durations of the 10 full transits (excluding 20080418) we fit the data using two models:
(1) a flat line, corresponding to a constant duration, with reduced χ2 = 0.9, and a value of
9636 ± 58 s, very similar to the joint-transit fit value, and (2) a sloped line, with reduced
χ2 = 0.6, and a slope of −0.24 ± 0.12 seconds per day. This is a much shallower slope than
the hint of a trend reported by D´ıaz et al. (2008), and is not significant given the good fit
achieved with the flat line.
Because the duration is a derived quantity, we must examine the parameters on which it
depends (i, a/R∗, and k). Both the inclination and the semi-major axis exhibit slight slopes,
but with low significance given the errors (i changes by 0.22 ± 0.09 degrees per year with
reduced χ2 = 0.2, and a/R∗ by 0.0007± 0.0002 stellar radii per year with reduced χ2 = 0.2).
If the duration were really decreasing because the planet is precessing, this would be due to
the planet moving away from the center of the stellar disk and hence the inclination would
decrease; instead, the best-fit slope for the inclination is slightly positive, another indication
that we are not picking up on a real effect.
On the other hand, the radius ratio does have real variations between transits, given the
current photometry. We find variation from a low of k = 0.118 ± 0.002 (transit 20061219,
from D´ıaz et al. 2008) to a high of k = 0.132 ± 0.002 (20060221 and 20060305, both from
Winn et al. 2007), with the rest of the transits in between (see Table 6). (Note that our
individual fits for the radius and error of the two transits from Winn et al. (2007) agree with
those cited in that paper, indicating that our fitting methods are comparable.) The best fit
line with a slope is not statistically significant (within 1σ of 0), and the most likely expla-
nation of the variation in radius depths is due to systematic effects in how the photometry
was reduced (see § 3.3).
If the star were active, the presence of stellar spots or active regions can affect the
-- 14 --
observed radius depths (see e.g. Pont et al. (2008) for HD189733, Rabus et al. (2009) for
TrES-1, and Czesla et al. (2009); Huber et al. (2009); Silva-Valio et al. (2010) for CoRoT-2,
all of which are known or theorized to be spotty stars). There is no record of variability
for this star in the current literature. We examined two published data sets with observa-
tions over multiple non-transit nights: 4 nights of VIMOS data (Pietrukowicz et al. 2010;
Minniti et al. 2007) and data spanning 115 nights from the OGLE survey Udalski et al.
(2002). We found that the long-term flux was stable to within a few mmag in both datasets
(3 mmag and 5 mmag respectively).
Since the trends in the parameters i, a/R∗, and k are slight and unlikely to be physical,
we cannot conclude based on the available data that the observed duration variation is a
real effect.
4.3. Limits on perturber mass
Although there is no clear evidence of TTVs beyond the 2σ level in the current dataset,
we can use the TTVs reported in Table 7 and shown in Figure 4 to place upper limits on
the mass and orbital separation of a hypothetical perturbing planet in the system. For that
purpose we use an implementation of the algorithm presented in Steffen and Agol (2005),
kindly provided by D. Fabrycky.
We explored the full perturber's mass parameter space for interior orbits and exterior
exterior orbits for orbital periods from 0.9-17.5 days (0.2-4.4 times the orbital period of
OGLE-TR-111b), and small initial eccentricity ec = 0.05 (we also examined ec = 0 and
ec = 0.3). All the orbits were assumed coplanar and the orbital instability regions in each
case were determined following Barnes and Greenberg (2006). The orbital period of the
perturber was increased by a factor of 1.1 for each step. For each period, the mass of the
perturber also increased from an initial value of 0.0001 M⊕ until reaching a mass that would
produce a TTV equivalent to the 3σ confidence level of our results (i.e. ∆χ2 = 9). We
used the time reported by Pietrukowicz et al. (2010) for the transit 20050409 instead of the
value we fit from the original photometry from Minniti et al. (2007), since the revised value
is more in line with expectations; simulations run omitting that transit yield similar results.
The mass limits placed by these tests are illustrated in Figure 4.
The constraints placed on the perturber's mass are strongest near the low-order mean
motion resonances, particularly in the interior and exterior 2:1 resonances, where we are
sensitive to objects as small as 1 M⊕ and 0.6 M⊕, respectively if ec = 0.05 (ec = 0 yields
even smaller constraints, but we choose to cite the more conservative value). No meaningful
-- 15 --
constraint can be placed on the region of the 4:1 external resonance, which was identified by
D´ıaz et al. (2008) as a possible location for a 1 M⊕ perturber that could explain their TTV,
but since we do not reproduce their TTV such a perturber is no longer necessary.
Finally, our O-C data, with a 1σ precision of 36-114 seconds (after accounting for system-
atic errors), cannot constrain the presence of moons around OGLE-TR-111b, which would
introduce a TTV signal of 4.6 seconds for a 1 M⊕ moon (Kipping et al. 2009).
5. Conclusions
We have tested the previously claimed presence of TTVs and TDVs in the OGLE-TR-
111 system by adding six new transit epochs, observed between 2008 and 2009, to the five
previously published results by Winn et al. (2007); Minniti et al. (2007); D´ıaz et al. (2008).
This new analysis not only doubles the number of available data points, but also extends
the TTV baseline from two to four years. In addition, combining the six new transits data
allows us to provide a new, more precise value of the radius of this planet. We find a new
radius for the planet of 1.019 ± 0.026 RJ , which is intermediate to the previously reported
radii by Winn et al. (2007) and D´ıaz et al. (2008), and is more precise.
We find a slight variation over time of the duration of the transits of OGLE-TR-111b,
as well as variations of other parameters, such as the inclination and semimajor axis of the
orbit. Those variations could, in principle, be attributed to perturbations of the orbit of
OGLE-TR-111b produced through interaction with additional planet(s) in the system, but
we demonstrate that the variations can be instead explained by systematic errors in the
data, and therefore should not be attributed to other planets.
We have also computed the transit midtimes of our new transits with formal precisions
of 20-40 seconds, and more accurate precisions of 35-50 seconds for the full transits (and
almost 2 minutes for the half-transit) once systematic errors are considered. The errors
on the literature transits similarly increased when the photometry is refit using the same
method to account for systematics, to 60-110 seconds depending on the light curve.
A longer time baseline and more precise timing data is still necessary to test further for
the presence of other planets in the OGLE-TR-111 system, especially in potentially stable
non-resonant orbits, but with the present results we conclude that OGLE-TR-111 belongs
in the category of systems summarized in Table 1 for which there is no sign of additional
planets more massive than a few M⊕ in low-order resonant orbits, including a limit of 1 M⊕
near the 2:1 resonances. The presence of massive (Earth-like) moons around OGLE-TR-111b
is still possible, but to detect those we would require timing precision of a few seconds or
-- 16 --
better, beyond the current capability of ground-based instrumentation for this system.
E.R.A. received support from NASA Origins grant NNX07AN63G. M.L.M. acknowl-
edges support from NASA through Hubble Fellowship grant HF-01210.01-A/HF-51233.01
awarded by the STScI, which is operated by the AURA, Inc.
for NASA, under contract
NAS5-26555. We would like to thank Brian Taylor and Paul Schechter for their tireless
instrument support. We thank the Magellan staff, in particular telescope operators Jorge
Araya, Mauricio Martinez, Hern`an Nunez, Hugo Rivera, Geraldo Valladares, and Sergio
Vera, for making these observations possible. We also thank Josh Carter and Dan Fabrycky
for providing software, and Josh Winn for some helpful discussions.
-- 17 --
REFERENCES
Agol, E., Steffen, J., Sari, R., and Clarkson, W.: 2005, MNRAS 359, 567
Agol, E. and Steffen, J. H.: 2007, MNRAS 374, 941
Ballard, S., Christiansen, J. L., Charbonneau, D., Deming, D., Holman, M. J., Fabrycky,
D., A'Hearn, M. F., Wellnitz, D. D., Barry, R. K., Kuchner, M. J., Livengood, T. A.,
Hewagama, T., Sunshine, J. M., Hampton, D. L., Lisse, C. M., Seager, S., and Vev-
erka, J. F.: 2009, ArXiv e-prints
Barnes, R. and Greenberg, R.: 2006, ApJ 647, L163
Bean, J. L.: 2009, A&A 506, 369
Bean, J. L. and Seifahrt, A.: 2008, A&A 487, L25
Carter, J. A. and Winn, J. N.: 2009, ApJ 704, 51
Claret, A.: 2000, A&A 363, 1081
Claret, A.: 2004, A&A 428, 1001
Coughlin, J. L., Stringfellow, G. S., Becker, A. C., L´opez-Morales, M., Mezzalira, F., and
Krajci, T.: 2008, ApJ 689, L149
Csizmadia, S., Renner, S., Barge, P., Agol, E., Aigrain, S., Alonso, R., Almenara, J. M.,
Bonomo, A. S., Borde, P., Bouchy, F., Cabrera, J., Deeg, H. J., De la Reza, R.,
Deleuil, M., Dvorak, R., Erikson, A., Guenther, E. W., Fridlund, M., Gondoin, P.,
Guillot, T., Hatzes, A., Jorda, L., Lammer, H., L´azaro, C., Leger, A., Llebaria, A.,
Magain, P., Moutou, C., Ollivier, M., Paetzold, M., Queloz, D., Rauer, H., Rouan,
D., Schneider, J., Wuchterl, G., and Gandolfi, D.: 2009, ArXiv e-prints
Czesla, S., Huber, K. F., Wolter, U., Schroter, S., and Schmitt, J. H. M. M.: 2009, A&A
505, 1277
D´ıaz, R. F., Rojo, P., Melita, M., Hoyer, S., Minniti, D., Mauas, P. J. D., and Ru´ız, M. T.:
2008, ApJ 682, L49
Dunham, E. W., Elliot, J. L., Bida, T. A., and Taylor, B. W.: 2004, in A. F. M. Moorwood
& M. Iye (ed.), Society of Photo-Optical Instrumentation Engineers (SPIE) Confer-
ence Series, Vol. 5492 of Presented at the Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference, pp 592 -- 603
-- 18 --
Ford, E. B. and Holman, M. J.: 2007, ApJ 664, L51
Gallardo, J., Minniti, D., Valls-Gabaud, D., and Rejkuba, M.: 2005, A&A 431, 707
Gelman, A. and Rubin, D. B.: 1992, Statistical Science 7, 457
Gibson, N. P., Pollacco, D., Simpson, E. K., Barros, S., Joshi, Y. C., Todd, I., Keenan, F. P.,
Skillen, I., Benn, C., Christian, D., Hrudkov´a, M., and Steele, I. A.: 2009a, ApJ 700,
1078
Gibson, N. P., Pollacco, D. L., Barros, S., Benn, C., Christian, D., Hrudkov´a, M., Joshi,
Y. C., Keenan, F. P., Simpson, E. K., Skillen, I., Steele, I. A., and Todd, I.: 2009b,
MNRAS pp 1684 -- +
Gillon, M., Pont, F., Moutou, C., Santos, N. C., Bouchy, F., Hartman, J. D., Mayor, M.,
Melo, C., Queloz, D., Udry, S., and Magain, P.: 2007, A&A 466, 743
Heyl, J. S. and Gladman, B. J.: 2007, MNRAS 377, 1511
Holman, M. J. and Murray, N. W.: 2005, Science 307, 1288
Huber, K. F., Czesla, S., Wolter, U., and Schmitt, J. H. M. M.: 2009, A&A 508, 901
Kipping, D. M.: 2009, MNRAS 396, 1797
Kipping, D. M., Fossey, S. J., and Campanella, G.: 2009, MNRAS 400, 398
Mandel, K. and Agol, E.: 2002, ApJ 580, L171
Miller-Ricci, E., Rowe, J. F., Sasselov, D., Matthews, J. M., Guenther, D. B., Kuschnig, R.,
Moffat, A. F. J., Rucinski, S. M., Walker, G. A. H., and Weiss, W. W.: 2008a, ApJ
682, 586
Miller-Ricci, E., Rowe, J. F., Sasselov, D., Matthews, J. M., Kuschnig, R., Croll, B., Guen-
ther, D. B., Moffat, A. F. J., Rucinski, S. M., Walker, G. A. H., and Weiss, W. W.:
2008b, ApJ 682, 593
Minniti, D., Fern´andez, J. M., D´ıaz, R. F., Udalski, A., Pietrzynski, G., Gieren, W., Rojo,
P., Ru´ız, M. T., and Zoccali, M.: 2007, ApJ 660, 858
Miralda-Escud´e, J.: 2002, ApJ 564, 1019
Osip, D. J., Floyd, D., and Covarrubias, R.: 2008,
in Society of Photo-Optical Instrumen-
tation Engineers (SPIE) Conference Series, Vol. 7014 of Presented at the Society of
Photo-Optical Instrumentation Engineers (SPIE) Conference, p. 9
-- 19 --
Pietrukowicz, P., Minniti, D., D´ıaz, R. F., Fern´andez, J. M., Zoccali, M., Gieren, W.,
Pietrzy´nski, G., Ru´ız, M. T., Udalski, A., Szeifert, T., and Hempel, M.: 2010, A&A
509(26), A260000+
Pont, F., Bouchy, F., Queloz, D., Santos, N. C., Melo, C., Mayor, M., and Udry, S.: 2004,
A&A 426, L15
Pont, F., Knutson, H., Gilliland, R. L., Moutou, C., and Charbonneau, D.: 2008, MNRAS
385, 109
Pont, F., Zucker, S., and Queloz, D.: 2006, MNRAS 373, 231
Rabus, M., Deeg, H. J., Alonso, R., Belmonte, J. A., and Almenara, J. M.: 2009, A&A 508,
1011
Santos, N. C., Ecuvillon, A., Israelian, G., Mayor, M., Melo, C., Queloz, D., Udry, S.,
Ribeiro, J. P., and Jorge, S.: 2006, A&A 458, 997
Silva, A. V. R. and Cruz, P. C.: 2006, ApJ 642, 488
Silva-Valio, A., Lanza, A. F., Alonso, R., and Barge, P.: 2010, A&A 510(26), A260000+
Simon, A., Szatm´ary, K., and Szab´o, G. M.: 2007, A&A 470, 727
Southworth, J.: 2008, MNRAS 386, 1644
Steffen, J. H. and Agol, E.: 2005, MNRAS 364, L96
Taylor, B. W., Dunham, E. W., and Elliot, J. L.: 2004,
in H. Lewis & G. Raffi (ed.),
Society of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, Vol.
5496 of Presented at the Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference, pp 446 -- 454
Udalski, A., Szewczyk, O., Zebrun, K., Pietrzynski, G., Szymanski, M., Kubiak, M., Soszyn-
ski, I., and Wyrzykowski, L.: 2002, Acta Astronomica 52, 317
Winn, J. N., Holman, M. J., and Fuentes, C. I.: 2007, AJ 133, 11
This preprint was prepared with the AAS LATEX macros v5.2.
1.00
0.95
0.90
0.85
0.80
0.75
20050409
20060221
20060305
20061219
20061223
20080418
20080422
20080512
20080516
20090217
20090313
-- 20 --
2.2 mmag2 min
1.6 mmag2 min
1.9 mmag2 min
2.2 mmag2 min
2.1 mmag2 min
1.3 mmag2 min
2.0 mmag2 min
1.5 mmag2 min
1.5 mmag2 min
1.2 mmag2 min
1.5 mmag2 min
-0.04
-0.02
0
Phase
0.02
0.04
Fig. 1. -- Eleven transits of OGLE-TR-111b. All available high-quality light curves are
plotted vs. orbital phase, with the data binned to 2 minutes to aid comparison. The joint
model fit (solid lines) were calculated using the parameter values in Table 5; the stated
standard deviation is the residuals from the joint model fit. Table 4 shows the unbinned
data; a full table is provided online.
0.0
0.1
0.2
0.3
0.4
u1,I
0.15
0.10
0.05
0.00
0.15
0.10
0.05
0.00
0.10
0.08
0.06
0.04
0.02
0.00
9400
9500
9600
9700
9800
9900
0.20
0.25
0.30
0.35
0.40
TC - 2453470 days
TC - 2453787 days
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.15
0.10
0.05
0.00
0.14
0.12
0.10
0.08
0.06
0.04
0.02
0.00
0.12
0.10
0.08
0.06
0.04
0.02
0.00
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.15
0.10
0.05
0.00
0.25
0.20
0.15
0.10
0.05
0.00
k
-- 21 --
aR*
i deg
0.10
0.08
0.06
0.04
0.02
0.00
0.15
0.10
0.05
0.00
0.122
0.124
0.126
0.128
0.130
11.5
12.0
12.5
13.0
87.0
87.5
88.0
88.5
89.0
89.5
90.0
b
T14 sec
u1,i'
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.561 0.562 0.563 0.564 0.565 0.566 0.567
0.706
0.707
0.708
0.709
0.710
0.711
TC - 2453799 days
TC - 2454092 days
TC - 2454088 days
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.748 0.749 0.750 0.751 0.752 0.753 0.754 0.755
0.788
0.790
0.792
0.794
0.796
0.800
0.802
0.804
0.806
0.808
TC - 2454574 days
TC - 2454578 days
TC - 2454598 days
0.540
0.541
0.542
0.543
0.544
TC - 2454602 days
0.15
0.10
0.05
0.00
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.25
0.20
0.15
0.10
0.05
0.00
0.552
0.553
0.554
0.555
0.556
0.625
0.626
0.627
0.628
0.629
TC - 2454879 days
TC - 2454903 days
0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
0.639
0.640
0.641
0.642
0.643
0.636
0.637
0.638
0.639
0.640
0.722
0.723
0.724
0.725
0.726
0.727
0.728
Fig. 2. -- Smoothed histogram of normalized parameter distributions, from which the pa-
rameters in Table 5 are derived. The solid line is the median value (which is very close to
the mean value in all cases). Note that several of the distributions, particularly a/R∗, i, and
b, are not strictly Gaussian. The dashed lines show the 68.3% credible interval. These values
were calculated for 2,850,000 links.
-- 22 --
ae
ae
ae
ae
ae
ae
g
e
d
i
ae
ae
90.0
89.5
89.0
88.5
88.0
ae
87.5
87.0
86.5
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
0.135
0.130
ae
0.125
k
ae
ae
0.120
0.115
0.110
0
200
400
600
800
1000
1200
1400
0
200
400
600
800
1000
1200
1400
days
days
c
e
s
4
1
T
10 200
10 000
ae
9800
9600
9400
9200
ae
ae
ae
ae
ae
ae
ae
*
R
a
ae
ae
0
200
400
600
800
1000
1200
1400
days
14
13
12
11
ae
0
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
200
400
600
800
1000
1200
1400
days
Fig. 3. -- Parameter variation vs. time for all transits of OGLE-TR-111b, based on individual
MCMC fits (Table 6), for (clockwise, from top-left): k, i, a/R∗, and T14. The errors have
been scaled upward based on the factor calculated from residual permutation. The values
derived from the joint MCMC fit to all transits (Table 5) are plotted as solid black lines with
±1σ errors. The dashed red lines indicate the best sloped fit with ±1σ errors, although all
fits are only marginally significant (within 1σ of a constant value for the radius ratio and
within 2σ for the other parameters).
-- 23 --
300
200
100
0
-100
-200
-300
300
200
100
0
-100
-200
-300
c
e
s
C
-
O
c
e
s
C
-
O
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
Diaz et al. 2008 ephemeris
-150
-100
-50
0
50
orbital periods
100
150
200
ae
ae
ae
ae
ae
ae
ae
ae
ae
ae
New ephemeris
-150
-100
-50
0
50
orbital periods
100
150
200
Fig. 4. -- Observed minus calculated midtimes for OGLE-TR-111b. Top panel: timing
residuals for eleven transits using the ephemeris from D´ıaz et al. (2008). Bottom panel:
timing residuals using the new ephemeris (Equation 2. The O-C values and errors are
shown in Table 7, and were calculated using the formal fit midtimes reported in Table 5
but rescaling the errors to more realistically account for systematic noise (see § 3.3). The
solid line represents zero deviation from expected time of transit, while the dashed lines
represent the 1σ and 3σ errors on the calculated orbital period, indicating the slopes that
result for a mis-determined period. We plot our calculated midtime for 20050409, based
on the photometry from Minniti et al. (2007), as an open circle, and the revised midtime
and error reported by Pietrukowicz et al. (2010) as a solid triangle. The new ephemeris was
calculated using only the solid symbols, i.e., using the Pietrukowicz et al. (2010) time.
-- 24 --
1:4
1:3
2:5
1:2
2:3
3:2
2:1
5:2
3:1
4:1
Å
M
1000
500
100
50
10
5
1
1.0
1.5
2.0
3.0
Period days
5.0
7.0
10.0
15.0
Fig. 5. -- Upper mass limit on potential companion planets vs. period of the perturber. We
examine companions with initial ec = 0.05(black) and ec = 0.0 (gray). The constraints are
strongest near the 2:1 mean-motion resonance, where objects as small as 1 M⊕ would have
been detectable; other interior and exterior resonances are also labeled. The shaded grey
region shows the instability region for a 1 M⊕, following Barnes and Greenberg (2006).
-- 25 --
Table 1. Maximum Allowed Perturber Mass (M⊕) Constrained by Transit Timing
System
Max O-C 1:2a
2:3
(sec)
Interior
(non-resonant)
3:2
2:1
Exterior Reference
(non resonant)
CoRoT-1
78 ± 24
< 60
unstable
--
unstable
--
unstable
--
GJ 436
HAT-P-3
HD 189733
HD 209458
< 60
< 120
< 60
< 45
70 ± 50
--
10
0.7-1.0
4
0.3
--
TrES-1
107 ± 17
1
TrES-2
TrES-3
257 ± 27
70 ± 30
2
3-4
--
2
--
1
--
--
2
15
--
--
--
--
--
20
--
3
30
--
--
--
--
2
--
8
0.3
--
2
--
--
4
1
--
20
30-40
20
0.3
--
1
1
10-15
200 (P < 4d)
Bean (2009)
10 (P < 3.5d), Csizmadia et al. (2009)
100 (P < 5d)
8 (P < 5d)
20 (P < 5d)
Bean and Seifahrt (2008)
Ballard et al. (2009)
-- Gibson et al. (2009b)
32 (P < 5d) Miller-Ricci et al. (2008b)
-- Miller-Ricci et al. (2008a)
17 (P < 7d), Agol and Steffen (2007)
100 (P < 10d)
100 (P < 6d)
Steffen and Agol
Rabus et al. (2009)
50 (P < 7d) Rabus et al. (2009)
-- Gibson et al. (2009a)
(2005);
aA perturber in an n:m resonance completes m orbits while the known planet completes n.
Table 2. Selected Parameters for OGLE-TR-111b
Radiusa
(RJ )
Midtime
(JD)
Duration
(hours)
Impact parameter
Filter Reference
24552330.44867 (fixed)
0.97 ± 0.06
24553470.56389 ± 0.00055
1.01 ± 0.06
24553787.70854 ± 0.00035
1.067 ± 0.054
1.067 ± 0.054
24553799.75138 ± 0.00032
0.922 ± 0.057b 24554088.79145 ± 0.00045
0.922 ± 0.057b 24554092.80493 ± 0.00045
--
2.9 ± 0.1
2.743 ± 0.033
2.743 ± 0.033
2.67 ± 0.014
2.67 ± 0.014
1.025 ± 0.047
1.066 ± 0.038
1.062 ± 0.051
0.952 ± 0.076
0.968 ± 0.064
1.045 ± 0.036
1.030 ± 0.029
0.996 ± 0.022
1.028 ± 0.021
1.006 ± 0.025
1.025 ± 0.022
24553470.56385 ± 0.00100
24553787.70860 ± 0.00076
24553799.75135 ± 0.00090
24554088.79118 ± 0.00138
24554092.80494 ± 0.00129
24554574.54805 ± 0.00078
24554578.55395 ± 0.00061
24554598.62680 ± 0.00047
24554602.64098 ± 0.00053
24554879.63787 ± 0.00056
24554903.72515 ± 0.00039
2.75 ± 0.11
2.73 ± 0.047
2.76 ± 0.065
2.62 ± 0.060
2.66 ± 0.071
2.40 ± 0.393c
2.73 ± 0.062
2.63 ± 0.037
2.69 ± 0.051
2.68 ± 0.054
2.61 ± 0.058
0 − 0.68
0 − 0.65
0.25 − 0.55
0.25 − 0.55
0.38 ± 0.2
0.38 ± 0.2
0.39 ± 0.28
0.30 ± 0.14
0.39 ± 0.16
0.38 ± 0.12
0.35 ± 0.15
0.51 ± 0.27
0.36 ± 0.19
0.22 ± 0.11
0.30 ± 0.12
0.28 ± 0.12
0.16 ± 0.11
aExcept as noted, assuming R∗ = 0.83R⊙
bUsing R∗ = 0.811R⊙; if R∗ = 0.83R⊙ is used, the radius is RP = 0.944RJ
cHalf-transit
I
V
I
I
I
I
V
I
I
I
I
i'
i'
i'
i'
i'
i'
Pont et al. (2004); Santos et al. (2006)
Minniti et al. (2007); D´ıaz et al. (2008)
Winn et al. (2007)
Winn et al. (2007)
D´ıaz et al. (2008)
D´ıaz et al. (2008)
this workd
this workd
this workd
this workd
this workd
this workd
this workd
this workd
this workd
this workd
this workd
dAll values from this work are taken from individual MCMC fits to each light curve, with with errors inflated to account for
correlated noise; see § 3.3.
-- 26 --
Table 3. Observational and Aperture Photometry Parameters for Six New Transits
Transit
(UT)
Frames
used (discarded)
20080418
20080422
20080512
20080516
20090217
20090313
178 (0)
110 (21c )
241 (20d )
276 (4e )
800 (0)
600 (42g )
aRadius around star.
Exp. Time Binning Readout NComp Aperturea Sky radius
(sec)
(pixels)
(sec)
(pixels)
Sky width
(pixels)
Scatterb Est. Poisson
(mmag)
(mmag)
30-60
120
60
30-100
30
15-30
1x1
1x1
1x1
1x1
2x2
2x2
5
5
5
5
0.003f
0.003f
3
4
3
7
6
1
12.8
16.4
19.2
17.4
19.2
9.6
35
30
30
30
25
20
15
10
15
10
20
30
1.3
2.0
1.5
1.5
1.2
1.5
1.1
0.8
0.8
0.8
1.1
1.5
bStandard deviation of the residuals on data binned to 120 s.
cInsufficient counts on target.
dElongated images due to tracking failure.
eInitial telescope focus not yet settled (3 points) and strongly aberrant ratio (1 point).
f Frame transfer mode.
gComparison star saturated.
-- 27 --
Table 4. Flux Values For New Transits of OGLE-TR-111b a
Mid-exposure (UTC) Mid-exposure (BJD)
Flux
Error
2454574.520046
2454574.522041
2454574.522868
2454574.523734
2454574.524134
· · ·
2454574.52375
2454574.525745
2454574.526573
2454574.527439
2454574.527838
0.9820188
0.9797899
0.9824249
0.9802172
0.9828871
0.00153
0.00153
0.00153
0.00153
0.00153
aFull table available online.
Table 5. Transit Parameters Based on Joint Fit to Eleven Light Curves
Median value
Formal Errora
Adopted Error b
0.1261
+0.0008, -0.0009
+0.0010, -0.0011
Fitted Parameters
k
a/R∗
i
u1,i′
u2,i′
u1,I
u2,I
u1,V
u2,V
TC − 2453470
TC − 2453787
TC − 2453799
TC − 2454088
TC − 2454092
TC − 2454574
TC − 2454578
TC − 2454598
TC − 2454602
TC − 2454879
TC − 2454903
Derived Parameters
b
T14 (sec)
Rp (RJ )c
a (AU)c
12.3
88.3
0.32
0.252
0.30
0.2582
0.6228
0.1587
0.56486
0.70934
0.75212
0.79197
0.80573
0.54282
0.55469
0.62752
0.64170
0.63864
0.72565
0.35
9647
1.019
0.0473
0.2
+0.3, -0.2
0.03
(fixed)
0.04
(fixed)
(fixed)
(fixed)
0.2
+0.3, -0.2
0.04
· · ·
0.05
· · ·
· · ·
· · ·
+0.00065,-0.00067 +0.00098, -0.00089
+0.00043,-0.00044 +0.00131, -0.00137
+0.00117, -0.00129
0.00044
0.00028
0.00032
0.00029
0.00025
0.00022
0.00025
+0.00071, -0.00077
+0.00080, -0.00097
+0.00040,-0.00039 +0.00103, -0.00068
+0.00044,-0.00045 +0.00061, -0.00061
+0.00055, -0.00047
+0.00044, -0.00046
+0.00044, -0.00035
+0.00029, -0.00042
+0.04, -0.06
+47, -48
0.025
+0.04, -0.06
+57, -60
0.026
+0.0015, -0.0014
+0.0015, -0.0014
aFormal 68.3% credible interval from MCMC fit of all 11 light curves jointly.
bAdopted error from residual permutation method, if greater than the formal error;
see § 3.3.
cAssuming R∗ = 0.83 ± 0.02 R⊙ (Santos et al. 2006) and using RJ = 71, 492km.
-- 28 --
Table 6. Transit Parameters Based on Individual Fits to Each Light Curve
Transit
20050409
20060221
20060305
20061219
20061223
20080418
20080422
20080512
20080516
20090217
20090313
ka
0.127 ± 0.0047
0.132 ± 0.0038
0.132 ± 0.0051
0.118 ± 0.0076
0.120 ± 0.0064
0.129 ± 0.0036
0.128 ± 0.0029
0.123 ± 0.0022
0.127 ± 0.0021
0.125 ± 0.0025
0.127 ± 0.0022
f b
1.0
2.0
3.2
3.1
1.8
1.2
1.2
1.9
1.5
2.2
1.9
a
T14
9901 ± 381
9829 ± 170
9920 ± 234
9436 ± 217
9580 ± 255
8646 ± 1413c
9837 ± 222
9482 ± 134
9677 ± 183
9644 ± 193
9390 ± 209
f b
· · ·
1.2
1.6
1.0
1.2
3.2
· · ·
1.5
1.4
1.9
2.7
a/R∗
a
f b
ia
11.7 ± 1.2
12.2 ± 0.7
11.8 ± 0.9
12.2 ± 0.8
12.2 ± 0.9
12.9 ± 1.4
12.9 ± 0.9
12.8 ± 0.6
12.3 ± 0.7
12.4 ± 0.7
13.0 ± 0.6
1.1
1.4
1.6
· · ·
1.1
1.8
1.2
1.5
1.3
1.6
2.2
88.1 ± 1.5
88.6 ± 0.8
88.1 ± 0.9
88.2 ± 0.9
88.3 ± 0.9
87.7 ± 1.5
88.3 ± 1.0
89.0 ± 0.6
88.6 ± 0.7
88.7 ± 0.7
89.2 ± 0.6
f b
1.3
1.1
1.4
· · ·
· · ·
2.5
1.1
1.0
· · ·
1.1
1.3
aFormal individual MCMC fit value and error (scaled upward by factor f in adjacent column).
bFactor by which the error in the previous column has been increased based on the residual
permutation method; no value is given if the formal MCMC fit error was larger. See § 3.3.
cThe ill-constrained duration of the half-transit 20080418 was not used in any fits.
Table 7. Timing Residuals
Transit
Number
O-C (s)
σ
20050409a
20050409b
20060221
20060305
20061219
20061223
20080418
20080422
20080512
20080516
20090217
20090313
−155
−155
−76
−73
−1
120
121
126
127
196
202
−271 ± 83
65 ± 43
8 ± 75
−41 ± 82
−66 ± 103
0 −125 ± 114
180 ± 87
−42 ± 52
10 ± 41
−13 ± 46
−3.3
1.5
0.1
−0.5
−0.6
−1.1
2.1
−0.8
0.3
−0.3
−1 ± 37 −0.02
28 ± 36
0.8
aUsing our analysis of the original photome-
try from Minniti et al. (2007).
bUsing the published time from reanalyzed
photometry by Pietrukowicz et al. (2010).
|
1902.01316 | 1 | 1902 | 2019-02-04T17:14:13 | A giant impact as the likely origin of different twins in the Kepler-107 exoplanet system | [
"astro-ph.EP",
"astro-ph.SR"
] | Measures of exoplanet bulk densities indicate that small exoplanets with radius less than 3 Earth radii ($R_\oplus$) range from low-density sub-Neptunes containing volatile elements to higher density rocky planets with Earth-like or iron-rich (Mercury-like) compositions. Such astonishing diversity in observed small exoplanet compositions may be the product of different initial conditions of the planet-formation process and/or different evolutionary paths that altered the planetary properties after formation. Planet evolution may be especially affected by either photoevaporative mass loss induced by high stellar X-ray and extreme ultraviolet (XUV) flux or giant impacts. Although there is some evidence for the former, there are no unambiguous findings so far about the occurrence of giant impacts in an exoplanet system. Here, we characterize the two innermost planets of the compact and near-resonant system Kepler-107. We show that they have nearly identical radii (about $1.5-1.6~R_\oplus$), but the outer planet Kepler-107c is more than twice as dense (about $12.6~\rm g\,cm^{-3}$) as the innermost Kepler-107b (about $5.3~\rm g\,cm^{-3}$). In consequence, Kepler-107c must have a larger iron core fraction than Kepler-107b. This imbalance cannot be explained by the stellar XUV irradiation, which would conversely make the more-irradiated and less-massive planet Kepler-107b denser than Kepler-107c. Instead, the dissimilar densities are consistent with a giant impact event on Kepler-107c that would have stripped off part of its silicate mantle. This hypothesis is supported by theoretical predictions from collisional mantle stripping, which match the mass and radius of Kepler-107c. | astro-ph.EP | astro-ph | A giant impact as the likely origin of different twins
in the Kepler-107 exoplanet system
Aldo S. Bonomo1*, Li Zeng2, Mario Damasso1, Zoë M. Leinhardt3, Anders B. Justesen4, Eric
Lopez5, Mikkel N. Lund4, Luca Malavolta6,7, Victor Silva Aguirre4, Lars A. Buchhave8, Enrico
Corsaro9, Thomas Denman3, Mercedes Lopez-Morales10, Sean M. Mills11, Annelies Mortier12, Ken
Rice13, Alessandro Sozzetti1, Andrew Vanderburg10,14, Laura Affer15, Torben Arentoft4, Mansour
Benbakoura16,17, François Bouchy18, Jørgen Christensen-Dalsgaard4, Andrew Collier Cameron12,
Rosario Cosentino19, Courtney D. Dressing20, Xavier Dumusque18, Pedro Figueira21,22, Aldo F. M.
Fiorenzano19, Rafael A. García16,17, Rasmus Handberg4, Avet Harutyunyan19, John A. Johnson10,
Hans Kjeldsen4, David W. Latham10, Christophe Lovis18, Mia S. Lundkvist4,23, Savita Mathur24,25,
Michel Mayor18, Giusi Micela15, Emilio Molinari26, Fatemeh Motalebi18, Valerio Nascimbeni6,7,
Chantanelle Nava10, Francesco Pepe18, David F. Phillips10, Giampaolo Piotto6,7, Ennio Poretti19,27,
Dimitar Sasselov10, Damien Ségransan18, Stéphane Udry18 & Chris Watson28
________________________________________________________________________________
1INAF -- Osservatorio Astrofisico di Torino, via Osservatorio 20, 10025 Pino Torinese, Italy; 2Department of Earth and
Planetary Sciences, Harvard University, Cambridge, MA 02138, USA; 3School of Physics, University of Bristol, HH
Wills Physics Laboratory, Tyndall Avenue, Bristol BS8 1TL, UK; 4Stellar Astrophysics Centre, Department of Physics
and Astronomy, Aarhus University, Ny Munkegade 120, DK-8000 Aarhus C, Denmark; 5NASA Goddard Space Flight
Center, 8800 Greenbelt Road, Greenbelt, MD 20771, USA; 6Dipartimento di Fisica e Astronomia "Galileo Galilei",
Università di Padova, Vicolo dell'Osservatorio 3, I-35122 Padova, Italy; 7INAF -- Osservatorio Astronomico di Padova,
Vicolo dell'Osservatorio 5, I-35122 Padova, Italy; 8DTU Space, National Space Institute, Technical University of
Denmark, Elektrovej 328, DK-2800 Kgs. Lyngby; 9INAF - Osservatorio Astrofisico di Catania, via S. Sofia 78, 95123,
Catania, Italy; 10Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA; 11The
Department of Astronomy, The California Institute of Technology, 1200 East California Blvd, Pasadena, CA 91125,
USA; 12Centre for Exoplanet Science, SUPA, School of Physics and Astronomy, University of St Andrews, St Andrews
KY16 9SS, UK; 13SUPA, Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill,
Edinburgh, EH93HJ, UK; 14Department of Astronomy, The University of Texas at Austin, 2515 Speedway, Stop
C1400, Austin, TX 78712, USA; 15INAF -- Osservatorio Astronomico di Palermo, Piazza del Parlamento 1, I-90124
Palermo, Italy; 16IRFU, CEA, Université Paris-Saclay, F-91191 Gif-sur-Yvette, France; 17Université Paris Diderot,
AIM, Sorbonne Paris Cité, CEA, CNRS, F-91191 Gif-sur-Yvette, France; 18Observatoire Astronomique de l'Université
de Genève, 51 ch. des Maillettes, 1290 Versoix, Switzerland; 19INAF -- Fundación Galileo Galilei, Rambla José Ana
Fernandez Pérez 7, E-38712 Breña Baja, Spain; 20Astronomy Department, University of California Berkeley, Berkeley,
CA 94720-3411, USA; 21European Southern Observatory, Alonso de Cordova 3107, Vitacura, Region Metropolitana,
Chile; 22Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, PT4150-762
Porto, Portugal; 23Zentrum für Astronomie der Universität Heidelberg, Landessternwarte, Königstuhl 12, 69117
Heidelberg, Germany; 24Departamento de Astrofísica, Universidad de La Laguna, E-38206, Tenerife, Spain; 25Instituto
de Astrofísica de Canarias, C/ Vía Láctea s/n, E-38205, La Laguna, Tenerife, Spain; 26INAF -- Osservatorio
Astronomico di Cagliari, Via della Scienza 5 -- I-09047 Selargius (CA), Italy; 27INAF -- Osservatorio Astronomico di
Brera, via E. Bianchi 46, 23807 Merate (LC), Italy; 28Astrophysics Research Centre, School of Mathematics and
Physics, Queen's University Belfast, Belfast BT7 1NN, UK. *e-mail: [email protected]
Measures of exoplanet bulk densities indicate that small exoplanets with radius less than 3
Earth radii ( R⊕ ) range from low-density sub-Neptunes containing volatile elements1 to higher
density rocky planets with Earth-like2 or iron-rich3 (Mercury-like) compositions. Such
astonishing diversity in observed small exoplanet compositions may be the product of
different initial conditions of the planet-formation process and/or different evolutionary paths
that altered the planetary properties after formation4. Planet evolution may be especially
affected by either photoevaporative mass loss induced by high stellar X-ray and extreme
ultraviolet (XUV) flux5 or giant impacts6. Although there is some evidence for the former7,8,
there are no unambiguous findings so far about the occurrence of giant impacts in an
exoplanet system. Here, we characterize the two innermost planets of the compact and near-
resonant system Kepler-107 (ref. 9). We show that they have nearly identical radii (about 1.5-
1.6 R⊕ ), but the outer planet Kepler-107c is more than twice as dense (about 12.6 g cm-3) as
the innermost Kepler-107b (about 5.3 g cm-3). In consequence, Kepler-107c must have a larger
iron core fraction than Kepler-107b. This imbalance cannot be explained by the stellar XUV
irradiation, which would conversely make the more-irradiated and less-massive planet
Kepler-107b denser than Kepler-107c. Instead, the dissimilar densities are consistent with a
giant impact event on Kepler-107c that would have stripped off part of its silicate mantle.
This hypothesis is supported by theoretical predictions from collisional mantle stripping 10,
which match the mass and radius of Kepler-107c.
The Kepler space telescope discovered the four sub-Neptune-sized planets Kepler-107b, c, d and e
(ref. 9), which transit (that is, pass in front of) their host star and have orbital periods (P) of 3.180,
4.901, 7.958 and 14.749 d, respectively. The four planets form a compact near-resonant system with
period ratios close or almost equal to ratios of small integers, specifically Pc/Pb=1.541 (~3:2),
Pe/Pc=3.009 (~3:1), Pd/Pb=2.503 (~5:2). The near resonances imply that the Kepler-107 planets
likely formed further out and then migrated inward getting trapped into resonances during the
migration process11. The orbital eccentricities of all four planets were determined to be low,
consistent with zero12, as expected from both dynamical stability criteria and orbit circularization
times for Kepler-107b and c (see Methods). Despite the proximity to resonances, no significant
transit timing variations from gravitational planet-planet interactions could be detected for any
planet13, preventing the direct determination of planetary masses from the Kepler light curve alone.
To determine the planet masses, and hence their bulk densities, we observed Kepler-107 from 2014
June 21 to 2017 April 22 with the fibre-fed high-resolution (R=115,000) HARPS-N spectrograph14
at the Telescopio Nazionale Galileo in La Palma. Our time series with 114 radial velocities has a
scatter of 6 m s-1 and shows two significant periodicities at 4.9 d and at about 14 d (see Methods and
Supplementary Figs. 5 and 6). The former is straightforwardly attributed to the Doppler signal
induced on the star by the second inner planet Kepler-107c; the latter is likely related to stellar
magnetic activity variations though at low level, as shown by the low amplitude of Kepler flux
variations (see Methods and Supplementary Fig. 2).
From the 114 HARPS-N spectra we determined the stellar atmospheric parameters (see Table 1 and
Methods). Even though the parent star is richer in iron than most of the stars that are known to host
low-mass planets15, it does not present an overabundance of iron relative to Si and Mg having solar-
like Fe/Si and Mg/Si photospheric ratios (see Methods and Supplementary Table 3). To determine
the stellar fundamental parameters (radius, mass and age) we performed a detailed asteroseismic
analysis with all the available Kepler short-cadence data. We extracted the frequencies for
individual modes of stellar oscillations from the power density spectrum of the Kepler light curve
(see Methods and Supplementary Fig. 4). We modelled them with the BAyesian STellar
Algorithm16 and found M ✶=1.238±0.029 M ⊙ , R✶=1.447±0.014 R⊙ , ρ✶=0.574±0.029 g cm-3,
logg=4.210±0.013 (cgs), in agreement with the spectroscopic value within 1σ, and an age
+0.70 Gyr (see Methods). These parameters are consistent with those that were previously
t= 4.29−0.56
determined17 but are more precise and accurate.
By using the constraint on the stellar density as provided by our asteroseismic analysis, we
modelled the transits of the four Kepler-107 planets (Supplementary Fig. 1) and derived their radii
and associated uncertainties by employing a differential evolution Markov chain Monte Carlo
Bayesian technique18 (see Table 1 and Methods). We confirm the absence of significant transit
timing variations (see Methods and Supplementary Fig. 8).
In the same Bayesian framework, we modelled the HARPS-N radial velocities with four Keplerians
by using the updated ephemerides from the transit fitting, and an additional sinusoid to account for
the activity signal at ~14 d (see Fig. 1 and Methods). More sophisticated models to take the activity-
induced correlated noise into account19 provided very consistent results (see Methods and
Supplementary Table 2). The radial velocity semi-amplitudes and derived planetary masses are
listed in Table 1. Only an upper limit could be established for the mass of Kepler-107d because its
Doppler signal is undetected due to its expected low mass and its orbital period longer than planets
b and c.
The positions of the Kepler-107 planets in the radius-mass diagram of small planets with a bulk
density measured better than 3σ are shown in Fig. 2 along with iso-composition20 curves. The two
innermost planets Kepler-107b and c have nearly equal radii, that is, Rp,b=1.536±0.025 R⊕ and
Rp,c=1.597±0.026 R⊕ , but dissimilar masses: Mp,b=3.51±1.52 M ⊕ and Mp,c=9.39±1.77 M ⊕ . This
means that they have different bulk densities with ρp,b=5.3±2.3 g cm-3 and ρp,c=12.6±2.4 g cm-3
(Table 1). According to theoretical models of planetary interiors20, both Kepler-107b and c are
consistent with a rocky composition. However, Kepler-107c appears Mercury-like in composition
and has an iron fraction that is at least a factor of two greater than that of Kepler-107b. Kepler-
107c's iron core makes up ~70% of its mass, with the silicate mantle making up ~30% (see
Methods). Kepler-107d could also be rocky, while Kepler-107e is a mini-Neptune containing a
significant fraction of volatiles in the form of a thin H/He envelope and/or water ice (see Fig. 2 and
Methods).
Unlike the pair of Kepler-36 neighbouring planets21, the difference in planetary density between
Kepler-107b and c cannot be explained by atmospheric escape (or photoevaporation) caused by the
high XUV stellar irradiation. This irradiation may partially or totally strip the gaseous H/He
envelopes of mini-Neptunes at close distances from the star and thus may yield a diversity in
exoplanet bulk densities7. However, it cannot be responsible for the observed dissimilar densities of
Kepler-107b and c because it would have led to a higher density for Kepler-107b than for Kepler-
107c (see Methods), which is at odds with our results.
Alternatively, the difference in density of the two inner planets can be explained by a giant impact
on Kepler-107c that removed part of its mantle, significantly reducing its fraction of silicates with
respect to an Earth-like composition22. The radius and mass of Kepler-107c, indeed, lie on the
empirically derived collisional mantle stripping curve for differentiated rocky/iron planets10,23 (grey
dashed line in Fig. 2). Smoothed particle hydrodynamics simulations show that a head-on high-
speed giant impact between two ~ 10 M ⊕ exoplanets in the disruption regime would result in a
planet like Kepler-107c with approximately the same mass and interior composition (see Fig. 3 and
Methods). Such an impact may destabilise the current resonant configuration of Kepler-107 and
thus it likely occurred before the system reached resonance. Multiple less-energetic collisions may
also lead to a similar outcome24.
Other planets with a Mercury-like composition such as K2-106b (ref. 25) and K2-229b (ref. 3) have
been recently discovered. However, they are all the innermost planets in their systems and thus
alternative mechanisms to the giant impact scenario may have given rise to their high iron content,
for instance mantle evaporation for the hottest planets3 or photophoresis and disc aerodynamic
fractionation yielding a depletion of silicates in the inner regions of the protoplanetary disc 26. These
other processes cannot explain the dissimilar densities of Kepler-107b and c (see Methods). An
alternative explanation might be that planet c formed closer to the parent star than planet b and
afterwards crossed its orbit; if so, the iron-rich composition of Kepler-107c might actually be
related to formation in the silicate-poor inner region of the protoplanetary disc26 with no need to
invoke a giant collision. However, this orbit crossing should have occurred before the dispersal of
the disc, so that the resulting eccentricities could have been damped. Given the relatively large mass
of planet c, it seems unlikely that there would have been sufficient time for such a scenario to
operate, or that it may have occurred without the eccentricities of the lower-mass planets becoming
large enough to destabilize the system.
Giant impacts are thought to have occurred in our Solar System and have been invoked to explain
the composition of Mercury27, the origin of the Earth-Moon system28 and the high orbital obliquity
of Uranus29. We have shown that they likely occurred in the exoplanetary system Kepler-107 and
shaped the compositional properties of its two inner planets. If catastrophic disruption impacts
occur frequently, we predict a clustering of small exoplanets along the maximum-collisional
stripping curve10 in the mass-radius diagram, as an increasing number of exoplanets are
characterized with precise radius and mass determinations.
Table 1: Properties of the Kepler-107 planetary system
Parameter
________________________________________________________________________________
Host star
Value and 68.3% credible interval
Kepler-107, KIC-10875245, KOI-117,
Magnitudesa
Distanceb (pc)
Systemic radial velocity Vr (km/s)
Effective Temperature Teff (K)
Metallicity [Fe/H] (dex)
Spectroscopic surface gravity logg (cgs)
Asteroseismic surface gravity logg (cgs)
Rotational velocity Vsini (km/s)
Mass M ✶ ( M⊙)
Radius R✶ ( R⊙)
Density ρ✶ (g cm-3)
Age t (Gyr)
2MASS 19480677+4812309
B=13.34, V=12.70, J=11.39, K=11.06
525.5 ± 5.5
5.64423 ± 4.5x10-4
5854 ± 61
0.321 ± 0.065
4.28 ± 0.10
4.210 ± 0.013
3.6 ± 0.5
1.238 ± 0.029
1.447 ± 0.014
0.574 ± 0.029
+0.70
4.29−0.56
Limb-darkening coefficients
________________________________________________________________________________________________
u1=0.26 ± 0.08, u2=0.56 ± 0.12
Planets
b
c
d
e
Orbital period P (d)
3.1800218 ± 2.9x10-6
4.901452 ± 1.0x10-5
7.95839 ± 1.2x10-4
14.749143 ± 1.9x10-5
Transit epoch
Tc - 2,450,000 (BJDTDB)
Transit duration (h)
Radius ratio R p/ R✶
Inclination i (deg)
Impact parameter b
Eccentricity e
Radial-velocity
semi-amplitude K (m s-1)
Radius R p ( R⊕)
Mass M p ( M ⊕)
Density ρ p (g cm-3)
Semi-major axis a (AU)
5701.08414 ± 3.7x10-4 5697.01829 ± 7.9x10-4
5702.9547 ± 0.0060
5694.48550 ± 4.6x10-4
3.633 ± 0.021
4.269 ± 0.037
4.18 ± 0.36
6.117 ± 0.031
0.00973 ± 1.3x10-4
0.01012 ± 1.3x10-4
0.00544 ± 3.6x10-4
0.01839 ± 1.4x10-4
89.05 ± 0.67
0.11 ± 0.08
0 (fixed)
1.32 ± 0.57
+0.34
89.49−0.44
+ 0.07
0.08− 0.05
0 (fixed)
3.06 ± 0.57
+0.64
87.55−0.48
+ 0.10
0.53− 0.14
0 (fixed)
< 1.06
89.67 ± 0.22
0.11 ± 0.07
0 (fixed)
1.95 ± 0.82
1.536 ± 0.025
1.597 ± 0.026
0.86 ± 0.06
2.903 ± 0.035
3.51 ± 1.52
5.3 ± 2.3
9.39 ± 1.77
12.65 ± 2.45
< 3.8
< 33.1
8.6 ± 3.6
2.00 ± 0.82
0.04544 ± 3.5x10-4
0.06064 ± 4.7x10-4
0.08377 ± 6.5x10-4
0.12639 ± 9.9x10-4
1593 ± 19
Equilibrium
temperaturec (K)
________________________________________________________________________________________________
(a): B and V Johnson magnitudes from the Howell Everett Survey; J and K magnitudes from 2MASS.
(b): From the second release Gaia data30.
(c): Black-body equilibrium temperature assuming a null Bond albedo and uniform heat redistribution to the night side.
1379 ± 17
1173 ± 14
955 ± 12
Fig. 1 Radial velocity measurements of Kepler-107 with HARPS-N. Unbinned (grey dots) and
binned (black circles) radial velocities with associated 1σ error bars are shown as a function of
orbital phase for the planets Kepler-107b, c and e (top left, top right, and bottom left panels). The
residuals after subtracting the Keplerian models (red solid lines) are also displayed. The radial
velocities of Kepler-107d are not shown because they exhibit no variations. The additional signal
with P~14 d, which we attribute to stellar activity, is shown on the bottom right panel.
Fig. 2 Mass -- radius diagram of exoplanets smaller than 3 R⊕ . Only known planets with bulk
densities measured to better than 33% precision through radial velocities (RV, circles) and transit
timing variations (TTV, triangles) are shown. Error bars represent 1σ confidence intervals. Colours
of symbols are a function of the planet irradiation level (or equilibrium temperature) assuming an
Earth-like albedo. The solid lines are theoretical mass-radius curves20 for planets with different
compositions as specified in the figure (see also Methods). The grey dashed line displays the
minimum radius predicted from collisional stripping models as a function of planetary mass10.
Fig. 3 Smoothed particle hydrodynamical collision simulation. Hemispheric cross-sectional
images of a high speed (62.5 km/s) head-on equal-mass ( 10.5 M ⊕ , 1.8 R⊕ ) simulation (see
Methods) are shown at four snapshots in time perpendicular to the impact plane. The strips display
composition (top), density (middle) and temperature (bottom). The impactors have identical initial
composition consisting of 70% forsterite mantle and 30% iron core by mass. The middle two frames
show a cross-sectional cut through the pancake structure created by the impact. The majority of the
material is vaporised during the impact. The largest post-collision remnant has a bound mass of
8.2 M ⊕ with a material composition of 36% forsterite and 64% iron, which is consistent with both
Kepler-107c mass and composition within 1σ. The post-collision remnant is significantly inflated
and still cooling at the end of the simulation (last frame).
References:
1. Lissauer, J. J. et al. All six planets known to orbit Kepler-11 have low densities. Astrophys. J.
770, 131 (2013).
2. Dressing, C. D. et al. The mass of Kepler-93b and the composition of terrestrial planets.
Astrophys. J. 800, 135 (2015).
3. Santerne, A. et al. An Earth-sized planet with a Mercury-like composition. Nature Astron. 2, 393-
400 (2018).
4 . Schlichting, H. E. Formation of Super-Earths. In Handbook of Exoplanets (eds Deeg, H. &
Belmonte, J.) 2345-2364 (Springer, Cham, 2018).
5. Lopez, E. D., Fortney, J. J. & Miller N. How Thermal Evolution and Mass-loss Sculpt
Populations of Super-Earths and Sub-Neptunes: Application to the Kepler-11 System and Beyond.
Astrophys. J. 761, 59 (2012).
6. Liu, S.-F., Hori, S., Lin D. N. C. & Asphaug E. Giant impact: an efficient mechanism for
devolatilization of super-earths. Astrophys. J. 812, 164 (2015).
7. Lopez, E. & Fortney, J. J. Understanding the Mass-Radius Relation for Sub-Neptunes: Radius as
a Proxy for Composition. Astrophys. J. 792, 17 (2014).
8. Lundkvist, M. S. et al. Hot super-Earths stripped by their host stars. Nature Communications 7,
11201 (2016).
9. Rowe, J. F. et al. Validation of Kepler's Multiple Planet Candidates. III. Light Curve Analysis
and Announcement of Hundreds of New Multi-planet Systems. Astrophys. J. 784, 45 (2014).
10. Marcus, R. A., Sasselov, D., Hernquist, L. & Stewart, S. T. Minimum radii of super-earths:
constraints from giant impacts. Astrophys. J. Lett. 712, L73-L76 (2010).
11. Terquem, C. & Papaloizou, J. C. B. Migration and the Formation of Systems of Hot Super-
Earths and Neptunes. Astrophys. J. 654, 1110-1120 (2007).
12. Van Eylen, V. & Albrecht, S. Eccentricity from Transit Photometry: Small Planets in Kepler
Multi-planet Systems Have Low Eccentricities. Astrophys. J. 808, 126 (2015).
13. Holczer, T. et al. Transit Timing Observations from Kepler. IX. Catalog of the Full Long-
cadence Data Set. Astrophys. J. Suppl. S. 225, 9 (2016).
14. Cosentino, R. et al. HARPS-N: the new planet hunter at TNG. Proc. SPIE 8446, 84461V
(2012).
15. Buchhave, L. A. et al. An abundance of small planets around stars with a wide range of
metallicities. Nature 486, 375-377 (2012).
16. Silva Aguirre, V. et al. Ages and fundamental properties of Kepler exoplanet host stars from
asteroseismology. Mon. Not. R. Astron. Soc. 452, 2127-2148 (2015).
17. Huber, D. et al. Fundamental Properties of Kepler Planet-candidate Host Stars using
Asteroseismology. Astrophys. J. 767, 127 (2013).
18. Bonomo, A. S. et al. Characterization of the planetary system Kepler-101 with HARPS-N. A
hot super-Neptune with an Earth-sized low-mass companion. Astron. Astrophys. 572, A2 (2014).
19. Damasso, M. et al. Eyes on K2-3: A system of three likely sub-Neptunes characterized with
HARPS-N and HARPS. Astron. Astrophys. 615, A69 (2018).
20. Zeng, L. & Sasselov, D. A detailed model grid for solid planets from 0.1 to 100 Earth masses.
Publ. Astron. Soc. Pac., 125, 227 (2013).
21. Lopez, E. & Fortney, J. J. The role of core mass in controlling evaporation: the Kepler radius
distribution and the Kepler-36 density dichotomy. Astrophys. J. 776, 2 (2013).
22. Leinhardt, Z. M. & Stewart, S. T. Collisions between Gravity-dominated Bodies. I. Outcome
Regimes and Scaling Laws. Astrophys. J. 745, 79 (2012).
23. Marcus, R. A., Stewart, S. T., Sasselov, D. & Hernquist, L. Collisional Stripping and Disruption
of Super-Earths. Astrophys. J. Lett. 700, L118-L122 (2009).
24. Chau, A., Reinhardt, C., Helled, R. & Stadel, J. Forming Mercury by Giant Impacts. Astrophys.
J. 865, 35 (2018).
25. Guenther, E. W et al. K2-106, a system containing a metal-rich planet and a planet of lower
density. Astron. Astrophys. 608, A93 (2017).
26. Ebel, D. S. & Stewart, S. T. The Elusive Origin of Mercury. In Mercury: The View After
MESSENGER (eds Solomon, S. C., Nittler, L. R. & Anderson, B. J.) 497-515 (Cambridge
University Press, Cambridge, 2018).
27. Benz, W., Anic, A., Horner, J. & Whitby, J. A. The Origin of Mercury. In Mercury (eds Balogh,
A. et al.) Space Sciences Series of ISSI, 7 -- 20, Vol. 26, (Springer, New York, 2008).
28. Lock, S. J. et al. The origin of the Moon within a terrestrial synestia. Journal of Geophys. Res.
123, 910-951 (2018)
29. Safronov, V. S. Sizes of the largest bodies falling onto the planets during their formation. Sov.
Astron. 9, 987 -- 991 (1966)
30. Berger, T. A. et al. Revised Radii of Kepler Stars and Planets using Gaia Data Release 2.
Astrophys. J. 866, 99 (2018).
Acknowledgments: The authors wish to thank Dr. R. D. Haywood, Dr. R. Silvotti and Prof. D. Charbonneau for useful
discussions. The HARPS-N project was funded by the Prodex Program of the Swiss Space Office (SSO), the Harvard-
University Origin of Life Initiative (HUOLI), the Scottish Universities Physics Alliance (SUPA), the University of
Geneva, the SmithsonianAstrophysical Observatory (SAO), the Italian National Astrophysical Institute (INAF),
University of St. Andrews, Queen's University Belfast and University of Edinburgh. The present work is based on
observations made with the Italian Telescopio Nazionale Galileo (TNG) operated on the island of La Palma by the
Fundación Galileo Galilei of the INAF (Istituto Nazionale di AstroFisica) at the Spanish Observatorio del Roque de los
Muchachos of the Instituto de Astrofísica de Canarias. This paper exploited data collected by the Kepler mission;
funding for the Kepler mission is provided by the NASA (National Aeronautics and Space Administration) Science
Mission directorate. The research leading to these results received funding from the European Union Seventh
Framework Programme (FP7/2007-2013) under grant agreement number 313014 (ETAEARTH). L.Z. acknowledges
support from the Simons Foundation (SCOL[award #337090]). MD acknowledges financial support from Progetto
Premiale 2015 FRONTIERA funding scheme of the Italian Ministry of Education, University, and Research. Funding
for the Stellar Astrophysics Centre is provided by The Danish National Research Foundation (Grant agreement No.
DNRF106). V.S.A. acknowledges support from VILLUM FONDEN (research grant 10118). E.C. is funded by the
European Union's Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement
No. 664931. T.D. is supported by a STFC PhD studentship. R.A.G. acknowledges support from CNES. This work has
been carried out in the frame of the National Centre for Competence in Research PlanetS supported by the Swiss
National Science Foundation (SNSF). C.L., F.B., F.P. and S.U. acknowledge the financial support of the SNSF. M.S.L.
is supported by The Independent Research Fund Denmark's Sapere Aude program (Grant agreement No. DFF -- 5051-
00130). S.M. acknowledges support from the Ramon y Cajal fellowship number RYC-2015-17697.
Author contributions: The underlying radial-velocity observation programme was conceived and organized by F.P.,
A.C.C., D.W.L., C.L., D. Ségransan, S.U. and E.M.. Observations with HARPS-N were carried out by L.A., A.C.C.,
A.M., C.D.D., M.D., X.D., R.C., A.F.M.F., P.F., A.H., F.M., M.L.M., L.M., C.N., V.N., K.R. and A.V.. C.L.
maintained and updated the reduction pipeline, L.M. implemented the correction of radial velocities for moonlight
contamination, and X.D. computed the values of the log(R'HK) activity indicator. A.S.B., M.D. and K.R. analysed and
modelled the radial velocities. L.M. simulated and compared the radial velocities using both non-interacting and
interacting Keplerians. A.S.B. also performed the transit fitting and S.M.M. worked on the analysis of transit timing
variations. A.S.B. and A.V. analysed the Kepler light curve in search of the stellar rotational modulation signal. L.B.
and A.M. determined the stellar atmospheric parameters from the HARPS-N spectra; A.M. also derived the stellar
chemical abundances. T.A., M.B., E.C., J.C.D., R.A.G., R.H., A.B.J., H.K., M.N.L., M.S.L., S.M., V.S.A. carried out
the asteroseismic analysis of the Kepler light curve and determined the stellar parameters from which A.S.B. and M.D.
derived the planetary orbital and physical parameters. L.Z. deduced the planet interior compositions and E.L. estimated
the planet atmospheric escapes. K.R. carried out dynamical simulations of the planetary system, and Z.M.L. and T.D.
conducted the simulations of giant impact events. A.S.B. was the primary author of the manuscript for which received
important contributions by L.Z., Z.M.L., M.D., E.L., M.N.L., V.S.A., A.S., A.M. and M.L.M.. All authors have
contributed to the interpretation of the data and the results.
Correspondence and requests for materials should be addressed to A.S.B.
METHODS
Kepler light curve, transit detection and stellar variability.
We downloaded the Kepler light curves extracted by both the Simple Aperture Photometry (SAP)
and Pre-search Data Conditioning (PDC) pipelines with long-cadence (29.42 min) and short-
cadence (58.8 s) sampling31. Photometric data were gathered over 4 years, from 2009 May 2 to 2013
May 11, in Kepler quarters (i.e., three-monthly intervals) 0-6, 8-10, 12-14, 16-17 in long cadence
and, starting from quarter 3, in short cadence as well. We performed a low-pass filter of the PDC
short-cadence light curve and searched for transits following ref. 32; we could thus confirm the
previously detected transit signals9 and found no additional transiting companions.
The long-cadence PDC light curve with transits removed shows a peak-to-peak amplitude of
~0.1%, with a maximum of ~0.2% between 1100 and 1300 BJDTDB-2,454,900 (Supplementary Fig.
2). We performed a weighted autocorrelation function of this light curve finding the highest, though
low-amplitude, peak at ~14 d (Supplementary Fig. 3). This peak might be the stellar rotation period
Prot, which would be consistent with the upper limit given by the projected rotational velocity Vsini
+14.5 d) at 1σ
(Table 1) assuming an edge-on stellar equator, that is, P rot<20.3−2.5
(3σ). No additional information on Prot could be extracted from the stellar activity indicators (see
below).
+3.3 d ( P rot<20.3−6.0
Asteroseismic analysis of Kepler data.
To perform the asteroseismic analysis on all the available short-cadence Kepler data, we first
corrected them using the KASOC33 and KADACS34 filter pipelines. We then computed the power
density spectrum from a weighted least-squares sine wave fitting, which was single-sided
calibrated, normalized according to Parseval's theorem, and converted to power density by dividing
by the integral of the spectral window35.
We performed a Bayesian peak-bagging analysis to extract the frequencies for individual modes of
oscillation, following the methodologies outlined in refs. 36-38. To define the priors for the mode
frequencies, we derived initial guesses by hand from smoothed versions of the power density
spectrum using the estimated value for the large frequency spacing ( Δ ν ), together with an
approximate value for the dimensionless offset ( ε ) for a proper mode identification39. Given the
low signal-to-noise ratio (S/N) of the oscillation power, we imposed Gaussian priors on the
oscillation mode linewidths based on the empirical relation for linewidth as a function of frequency
given in ref. 39; if uninformative priors are used at the S/N observed for this star, one simply risks
fitting noise spikes near the initial guess for the mode frequency. Following the procedure described
in ref. 40, we calculated for each fitted mode a metric for the quality of the fit which was adopted in
the stellar modelling efforts to decide on the number of modes to include in the analysis.
Supplementary Fig. 4 shows the power density spectrum of Kepler-107 together with the best fitting
model spectrum from the peak bagging.
Stellar parameters.
Our HARPS-N spectra were reduced with the online Data Reduction Software and used to
determine the stellar atmospheric parameters, that is Teff, logg, [Fe/H], with two different
techniques: the first one relies on the measure of the equivalent widths of a set of iron lines41, while
the second one compares the observed spectra with a library grid of synthetic template spectra 15.
The two methods provided fully consistent parameters and we took the averages of their values and
error bars as the final parameters and uncertainties (Table 1). The latter technique also allowed us to
estimate the stellar projected rotational velocity Vsini=3.6±0.5 km s-1. Moreover, we derived the
stellar abundances relative to the Sun (Supplementary Table 3) by using the same method as in ref.
42 and the solar abundances in ref. 43.
We determined the host star physical parameters, that is density, radius, mass and age (Table 1), by
fitting the peak-bagged oscillation frequencies to asteroseismic predictions from stellar evolution
models with the BAyesian STellar Algorithm16. Briefly, we constructed a fine grid of stellar tracks
using the GARching STellar Evolution Code44 and computed the associated theoretical frequencies
of oscillation with the Aarhus adiabatic oscillation package45. The adopted input physics includes
microscopic diffusion of helium and heavy elements following prescription and overshooting given
in ref. 46 and based on the model by ref. 47. To avoid the influence of the so-called surface effect in
our calculations48, we chose to reproduce ratios of oscillation frequencies that have been shown to
minimize the influence of the outermost stellar layers (see, for example, ref. 49 and references
therein). The derived stellar radius, R✶=1.447 ± 0.014 R⊙ agrees within 1σ with that determined
from the Gaia parallax30, that is R✶=1.45± 0.06 R⊙ , but is significantly more precise.
Radial-velocity observations and modelling.
In total we gathered 120 HARPS-N spectra by using typical exposure times of 30 min. The radial
velocities (RVs) were extracted from the HARPS-N spectra by performing a weighted cross
correlation with a numerical spectral mask of a G2V star50 and have a median uncertainty of 4.5
m s-1. Twenty-two RVs were contaminated by moonlight and were corrected with the method
described in ref. 51. Four measurements were discarded because of low S/N, which yields RV
uncertainties larger than 10 m/s; two more RV points were identified as outliers on the basis of
Chauvenet's criterion and also excluded. Our RV time series thus contains 114 RV measurements
that passed these criteria, and is shown in Supplementary Fig. 5. The RVs, their epochs and
uncertainties are listed in Supplementary Table 4 along with the measurements of activity indicators
such as the full-width at half maximum (FWHM) and the bisector (BIS) of the averaged line
profile52, which were computed from the cross-correlation function, and the logarithm of the R'HK
parameter defined as the ratio of the chromospheric emission in the cores of the CaII H and K lines
to the stellar bolometric emission53.
We searched for periodic signals in the RV and activity indicator time series by using Generalised
Lomb-Scargle (GLS) periodograms54 and taking the measurement uncertainties into account. The
obtained power spectra along with the false alarm probabilities54 and the spectral window are shown
in Supplementary Fig. 6. In the RV time series the most significant periodicities are found at
4.899±0.003 d and 13.79±0.03 d. The peak at 4.9 d corresponds to the orbital period of Kepler-107c
as determined from the observation of its transits in the Kepler light curve, and this RV signal is
indeed in phase with the orbital ephemeris of Kepler-107c (see top-right panel in Fig. 1) as we
expect if it is planetary in origin. The peak at 13.8 d is accompanied by a close peak at 14.30±0.03 d
which falls at the secondary peak of the spectral window; therefore, one of the two close peaks is
the alias of the other. Regardless of the ambiguity, this periodicity at ~14 d is likely related to stellar
activity because it does not correspond to any of the transit periods and is compatible with the
highest peak in the autocorrelation function of the Kepler light curve (Supplementary Fig. 3). Other
peaks that are just above the significance threshold in Supplementary Fig. 6 (top panel) disappear as
soon as the 4.9 d and 13.8 d signals are removed from the original dataset and are thus aliases as
well. No significant periodicities are found in the activity indicator measurements at the RV peaks
(Supplementary Fig. 6) and there are no significant correlations between activity indices and RVs.
In general, the host star is not active with a median value of log(R'HK) equal to -5.08±0.14 dex.
To determine the RV semi-amplitudes of the four Kepler planets and hence their masses, we
modelled the RVs with four non-interacting Keplerians, an additional sinusoid to take the activity
signal at ~14 d into account, the systemic radial velocity, and an RV uncorrelated jitter term to
account for extra noise as in ref. 55. We first adopted circular orbits for all the four planets; indeed,
the orbit circularization times of the innermost planets Kepler-107b and Kepler-107c due to tidal
+0.70 Gyr. Specifically, by
dissipation inside the planets56 are shorter than the system age, t= 4.29−0.56
' ≈1500 (ref. 57) which is reasonable for
assuming an Earth-like modified tidal quality factor Q p
rocky planets (and likely overestimated for Mercury-like planets), we find circularization times of
~0.13 and ~1.7 Gyr for planets b and c, respectively. We also adopted circular orbits for Kepler-
107d and Kepler-107e in the absence of any useful constraint on the orbital eccentricity from RVs,
given the non-detection of the Kepler-107d Doppler signal and the low-amplitude of that of Kepler-
107e. This is a reasonable assumption given that N-body dynamical simulations we ran for 10 Myr,
using the Mercury6 code58 with the masses shown in Table 1, indicate that the eccentricities of
planets d and e must be lower than 0.15 to achieve dynamical stability. Eccentricities consistent
with zero for all four planets were also found from Kepler photometry uniquely12.
The posterior distributions of the 17 free parameters were obtained in a Bayesian framework by
employing a differential evolution Markov chain Monte Carlo tool (DE-MCMC), following the
prescriptions given in refs. 18 and 59 and, in particular, running 34 chains (twice the number of free
parameters). Uninformative priors with reasonably large bounds were imposed on all the parameters
except for the transit mid-times and periods of the four planets for which Gaussian priors were set
from Kepler photometry (Supplementary Table 1). The medians and the 15.86% and 84.14%
quantiles of the posterior distributions were taken as the best values and 1σ uncertainties, and are
reported in Table 1 and Supplementary Table 2. Very consistent results on orbital parameters and
planetary masses are achieved when including the six discarded RVs and/or without performing the
moonlight correction.
We also fitted the RVs by employing a Gaussian process (GP) regression60 with a quasi-periodic
kernel to model the contribution due to the stellar variability following ref. 19. We used the
MultiNest v3.10 Bayesian inference tool61 with the same implementation as in ref. 62 to sample the
parameter space and derive the posterior distributions for the 18 model parameters. The latter are
the orbital parameters (period, transit time, RV semi-amplitude) of the four Kepler-107 planets, the
systemic RV, the uncorrelated jitter term and the four GP covariance hyper-parameters, that is, the
amplitude h of the correlations, the periodicity θ which can be related to the stellar rotation period,
the length scale w of the periodic component and the correlation decay time scale λ, which can be
associated with the lifetime of active regions (see equation 1 in ref. 19). The adopted prior
distributions for each fitted parameter are listed in Supplementary Table 1. We used 800 random
walkers (live points) and set the sampling efficiency parameter to 0.5. The RV semi-amplitudes and
their 1σ uncertainties obtained with both the GP analysis and the sinusoidal-activity model are fully
consistent (Supplementary Table 2). However, some of the GP hyper-parameters, such as the λ, w
and, to a lesser extent, h parameters are not well constrained. Nonetheless, we note that the θ
parameter is quite well confined, close to the expected ~14 d activity signal periodicity, even if the
used uniform prior interval was large enough, that is [12, 16] d.
Bayesian model comparison between the sinusoidal-activity model and that employing GP
regression was performed using MultiNest61 and the same priors as for the latter model (third
column in Supplementary Table 1). The obtained Bayes' factor of ~3.4 in favour of the sinusoidal-
activity model does not support the choice of the more complex GP model63.
To evaluate the possible effects of low but non-zero eccentricities on the recovered RV semi-
amplitudes and hence on the planetary masses, we re-did the DE-MCMC orbital fit by letting the
eccentricity of the four planets vary, though with reasonable priors. Specifically, we imposed
Gaussian priors with zero mean and standard deviation σe=0.05 on the eccentricities of Kepler-107b
and c, and Rayleigh priors with σe=0.04 for the two outer planets64. The derived RV semi-
amplitudes Kb=1.32±0.57, Kc=3.08±0.58, Kd<1.08 and Ke=1.89±0.83 m s- 1 are indistinguishable
from those obtained with circular orbits (Table 1).
As a further investigation, we checked that the use of non-interacting Keplerians is accurate enough.
To this end, we compared the planetary Doppler signals computed with both non-interacting and
interacting Keplerians using the TRADES code65. The maximum difference between the two
models is about 5 cm s-1 and is thus completely negligible.
The mass of Kepler-107c, Mp,c, has been determined with a relative uncertainty of 19% (better than
5σ), while that of Kepler-107b (Mp,b) with an uncertainty of 43% given that photon-noise RV error
bars are on average ~3.5 times higher than the RV semi-amplitude of planet b. According to these
uncertainties, the probability that Mp,b and Mp,c would differ by chance is 0.6% and 0.1% for the
sinusoidal-activity and GP models, respectively. To further show that Mp,b must be considerably
lower than Mp,c despite its low precision, we simulated artificial RV time series by (1) injecting at
the observation epochs the Doppler signal of Kepler-107b for three different mass values: Mp,b=Mp,c
( 9.4 M ⊕ ), Mp,b=0.5 Mp,c ( 4.7 M ⊕ ) and Mp,b=0.33 Mp,c ( 3.1 M ⊕ ); (2) adding the Doppler signals of
planets c and e as well as the 13.8 d activity signal with their semi-amplitudes as reported in Table 1
and Supplementary Table 2; and (3) shifting each RV point generated from the previous steps,
following a normal distribution with mean and standard deviation equal to the RV value and its 1σ
error bar. In this way, for every mass value of Kepler-107b we simulated 500 RV time series with
different noise realisations and computed their GLS periodogram. The averaged periodograms for
each simulated Mp,b value are displayed in Supplementary Fig. 7. These artificial periodograms
clearly show that Mp,b is lower than Mp,c by at least a factor of ~2, as found by the Bayesian
analyses, otherwise the power at Pb would have been greater than observed (top panel), and even
higher than the peak at Pc for Mp,b=Mp,c (second top panel).
Transit fitting and planetary parameters.
To perform the transit fitting, we treated the SAP short-cadence light curve as in ref. 66 and used
the transit model of ref. 67. For each planet we fitted for the transit epoch, the orbital period, the
transit duration, the radius ratio R p / R✶ , the orbital inclination, and the limb-darkening coefficients
q1=(u 1+u2)2 and q 2=0.5 u1 /(u 1+u 2) (ref. 68), where u1 and u2 are the coefficients of the limb-
darkening quadratic law69. We used circular orbits for all the planets as for the RV modelling. We
imposed a Gaussian prior on the stellar density following, for example, ref. 70, by taking advantage
of the very precise and accurate value derived from the asteroseismic analysis of Kepler data (
ρ✶=0.574±0.029 g cm-3 ). No bounds were set on the transit parameters except for the q1 and q2
parameters for which intervals of [0, 1] were adopted68.
The posterior distributions of the free parameters were obtained with the same DE-MCMC
technique as above and were used to derive the medians and 1σ confidence intervals, which are
reported in Table 1 except for q1 and q2 which were found to be 0.67 ± 0.10 and 0.56 ± 0.12,
respectively. The derived radii of planets b, c and e are fully consistent (within 1σ) with those
which were previously determined in ref. 12; for planet d we found a slightly smaller radius. The
planetary orbital and physical parameters were determined by combining the posterior distributions
of the stellar, RV and transit parameters, and are also listed in Table 1.
Moreover, we computed the transit timing variations (TTVs) for each planet with the short-cadence
data in the same Bayesian framework. Because of the low S/N of the individual transits, in
particular for planets b, c and d, we fitted to each transit the model obtained with the whole light
curve by letting only the mid-transit time vary. The Observed-Calculated (O-C) transit times of
planets b, c and e showing no significant variations from a constant ephemeris are displayed in
Supplementary Fig. 8. Those of planet d could not be determined because of the very low S/N of the
individual transits (see Supplementary Fig. 1, bottom-left panel), preventing the DE-MCMC chains
from achieving proper convergence in a single transit fitting, and are not shown.
Our non-detection of transit timing variations is in agreement with the results of ref. 13. Using the
formalism from ref. 71, one can estimate the expected TTV amplitudes due to neighbouring planets
with e≈0 . Planet b's TTV amplitude is expected to be ~3 min due to planet c; the TTV amplitude
of planet c is ~2 min from planet b and also ~2 min from planet d, using the RV measured (or upper
limit) masses. The interior planets have average TTV uncertainties of 25 min and hence the non-
detection of TTVs is not surprising. We also ran a photodynamic MCMC 72 to determine the upper
limits of the masses allowed with the observed transit data. We find upper limits of 22, 24, 8.6 and
22 M ⊕ for planets b, c, d and e at the 1σ level, which are all significantly higher than the masses
and upper limits measured via RVs (Table 1).
Planetary composition.
The equation-of-state data allow calculations of planet average densities under self-gravitational
compression and further constrain planet bulk compositions and internal structures following the
methods in ref. 20. The Earth-like composition is assumed to be 32.5 wt.% (weight percent) Fe/Ni-
metal plus 67.5 wt.% MgSiO3-rock, fully differentiated into a bi-layer core-mantle structure. 32.5%
+0.20 and Mg/Si= 1.22−0.21
of Fe/Ni-metal by mass is consistent with both the general cosmic element abundance (number)
ratio of Mg:Si:Fe being close to 1:1:1 and the Fe/Si and Mg/Si ratios derived from the host star
+0.25 , which can be
abundances (Supplementary Table 3), that is, Fe/Si= 0.89−0.17
considered as a proxy for the protoplanetary disc composition. According to these ratios, the iron-
rich composition of Kepler-107c cannot be primordial73. Moreover, Fe/Ni-metal and Mg-silicates
have similar volatility and condensation temperatures74 in nebula; therefore, it is rather difficult to
alter these ratios in the planet composition through merely temperature effects in the protoplanetary
disc, and in particular for the Kepler-107 system where the outer planet (c) is denser than the inner
one (b). A special scenario such as a giant impact must be invoked to account for that.
Together, Fe/Ni-metal and Mg-silicates comprise about half a percent of the nebula mass in a solar-
metallicity nebula. Differentiation is expected to occur on all rocky objects during their formation
due to melting or partial melting from the energy of early radioactive elements and accretion heat.
H2O is assumed to be in solid form along the melting curve (liquid-solid phase boundary). The 50%
H2O curve in Fig. 2 corresponds to an Earth-like composition with exactly equal mass of H 2O on
top. The ices that condensed out of the nebula are expected to be a mixture of H 2O-NH3-CH4. The
condensation of ices is expected to occur near icelines. If all H2O ice condenses out of the gas
phase, it is about equal mass as Fe/Ni-metal plus Mg-silicates, that is, another half a percent of the
uncondensed nebula by mass. And if NH3-CH4-clathrate-ices all condense out, it is yet another half
a percent of the uncondensed nebula by mass. The mass-radius curve of H 2O-NH3-CH4 mixture is
expected to be very similar to that of pure H2O . The H2-He gaseous content is assumed to be a
cosmic mixture of 75% H2 and 25% He by mass. Changing the metallicity of the host star and thus
of the nebula, given that the star Kepler-107 is metal-rich, will change the total amount of
condensed materials with respect to H2-He gas, but it will not change the relative mass ratio of Fe
metal versus Mg silicates, H2O ice, NH3-CH4-clathrate-ices, as this ratio is ultimately dictated by
the cosmic nuclear synthesis processes occurring in giant star interiors and supernovae over the
history of our Milky Way galaxy.
Planetary evolution and escape simulations.
Planetary evolution models were run using the coupled structure, thermal evolution and
photoevaporation model of ref. 75. As the inner planets in the system are all consistent with bare
rocky compositions, while the mass and radius of planet e require that it has a significant gaseous
envelope, for our simulations we assumed that these planets initially formed with solar composition
hydrogen and helium envelopes atop rock and iron cores. We then calculated the mass, radius and
composition evolution of each planet due to thermal cooling and atmospheric escape from XUV-
driven photoevaporative escape.
For a given planet composition and a given star, a planet's vulnerability to photoevaporative escape
−2 (ref. 21), where Fp is the incident bolometric flux that a planet receives
scales roughly as F p M core
from its parent star and Mcore is the planet core mass. As a result, the relatively low masses and high
irradiations of planets b and d mean that they are the most vulnerable to atmospheric escape,
followed by planet c. At ~8.6 M ⊕ and 14.75 d, however, planet e is massive enough and far enough
out that it is not nearly as vulnerable to photoevaporation, consistent with the fact that it appears to
have retained a gaseous envelope. Assuming an Earth-like core, we estimate that Kepler-107e
should have ~3% of its mass in a gaseous H/He envelope today and that only had ~5% when it
initially formed. On the other hand, due to its higher irradiation Kepler-107c could have easily lost a
~5% envelope in its first Gyr and planets b and d significantly more than that.
While photoevaporation may explain the overall density contrast between the inner three planets
and planet e, it cannot explain the contrast between planets b and c. Planet b is more than an order
of magnitude more vulnerable to photoevaporation than c, mostly because of its lower mass, and yet
it is planet c that appears to be denser and much more iron-rich. Additionally, we note that at just
~1400 K, Kepler-107c is below the vaporization temperature of most silicate species 76, suggesting
that vaporization and escape of the silicate mantle is unlikely to be responsible either. Indeed,
silicate atmospheres do not experience any significant escape until they reach temperatures >2000
K (ref. 77), hotter than any of the planets in this system. Moreover, while the exact scaling may be
different for other processes such as outgassing and escape from silicate atmospheres77 or minor
impact erosion78,79, all escape processes should be more effective for planet b than c, as it is both
more irradiated and less gravitationally bound. This qualitative fact alone suggests that the high
density of planet c is unlikely to be due to a gradual escape process that affected all planets in the
system consistently and must instead be due to a stochastic mechanism such as a giant impact.
Giant impact simulations.
A range of smoothed particle hydrodynamical simulations were run at various impact speeds using
a modified version of GADGET280. We focused on head-on and nearly head-on equal-mass
disruptive events because these are generally the most efficient at stripping outer layers 22. The
thermodynamic properties of the different materials were calculated at each step from a tabulated
equation of state23,81, and the initial thermodynamic profiles were determined following ref. 82.
Each initial impactor had a resolution of 105 particles. The post-collision bound mass was calculated
by taking the particle closest to the potential minimum as a seed, then iteratively calculating which
additional particles were gravitationally bound to the current set. This process continued until the
mass difference between iterations was below a 10-7 set tolerance83. We find that we can
significantly change the mantle-core ratio of the target as a result of a single high-energy impact, an
example of which is shown in Fig. 3. Such an impact could take a planet that originally was
dominated in mass by mantle (something like planet b) and convert it into a planet dominated in
mass by core (similar to planet c). Multiple energetic collisions may yield a similar outcome24 but,
given the large variety of possible configurations, their simulation goes beyond the scope of the
present work.
Data availability
The RV data that support the findings of this study and have been used to produce some of the plots
are available in the Supplementary Information. Kepler data are available at the Mikulski Archive
for Space Telescopes (https://archive.stsci.edu/kepler/).
References:
31. Jenkins, J. M. et al. Overview of the Kepler Science Processing Pipeline. Astrophys. J. Lett. 713,
L87-L91 (2010).
32. Bonomo, A. S. et al. Detection of Neptune-size planetary candidates with CoRoT data.
Comparison with the planet occurrence rate derived from Kepler. Astron. Astrophys. 547, A110
(2012).
33. Handberg, R. & Lund, M. N. Automated preparation of Kepler time series of planet hosts for
asteroseismic analysis. Mon. Not. R. Astron. Soc. 445, 2698-2709 (2014).
34. García, R. A. et al. Preparation of Kepler light curves for asteroseismic analyses. Mon. Not. R.
Astron. Soc. 414, L6-L10 (2011).
35. Kjeldsen, H. & Frandsen, S. High-precision time-resolved CCD photometry. Publ. Astron. Soc.
Pac. 104, 413 -- 434 (1992).
36. Handberg, R. & Campante, T. L. Bayesian peak-bagging of solar-like oscillators using MCMC:
a comprehensive guide. Astron. Astrophys. 527, A56 (2011).
37. Lund, M. N. et al. Standing on the Shoulders of Dwarfs: the Kepler Asteroseismic LEGACY
Sample. I. Oscillation Mode Parameters. Astrophys. J. 835, 172 (2017).
38. Corsaro, E. & De Ridder, J. DIAMONDS: a new Bayesian nested sampling tool. Application to
peak bagging of solar-like oscillations. Astron. Astrophys. 571, A71 (2014).
39. White, T. R. et al. Asteroseismic Diagrams from a Survey of Solar-like Oscillations with
Kepler. Astrophys. J. Lett. 742, L3 (2011).
40. Davies, G. R. et al. Oscillation frequencies for 35 Kepler solar-type planet-hosting stars using
Bayesian techniques and machine learning Mon. Not. R. Astron. Soc. 456, 2183-2195 (2016).
41. Mortier, A. et al. Correcting the spectroscopic surface gravity using transits and
asteroseismology. No significant effect on temperatures or metallicities with ARES and MOOG in
local thermodynamic equilibrium. Astron. Astrophys. 572, A95 (2014).
42. Mortier, A. et al. New and updated stellar parameters for 90 transit hosts. The effect of the
surface gravity. Astron. Astrophys. 558, A106 (2013).
43. Asplund, M., Grevesse, N., Sauval, A. J. & Scott, P. The Chemical Composition of the Sun.
Annu. Rev. Astron. Astr. 47, 481-522 (2009).
44. Weiss, A. & Schlattl, H. GARSTEC---the Garching Stellar Evolution Code. The direct
descendant of the legendary Kippenhahn code. Astrophys. Space Sci. 316, 99-106 (2008).
45. Christensen-Dalsgaard, J. ADIPLS---the Aarhus adiabatic oscillation package. Astroph. Space
Sci. 316, 113-120 (2008).
46. Thoul, A. A., Bahcall, J. N. & Loeb, A. Element diffusion in the solar interior. Astrophys. J. 421,
828-842 (1994).
47. Freytag, B., Ludwig, H.-G. & Steffen, M., Hydrodynamical models of stellar convection. The
role of overshoot in DA white dwarfs, A-type stars, and the Sun. Astron. Astrophys. 313, 497-516
(1996).
48. Kjeldsen, H., Bedding, T. R. & Christensen-Dalsgaard, J. Correcting Stellar Oscillation
Frequencies for Near-Surface Effects. Astrophys. J. Lett. 683, L175 (2008).
49. Silva Aguirre, V. et al. Stellar ages and convective cores in field main-sequence stars: first
asteroseismic application to two Kepler targets. Astrophys. J. 769, 141 (2013).
50. Pepe, F. et al. The CORALIE survey for southern extra-solar planets VII. Two short-period
Saturnian companions to HD 108147 and HD 168746. Astron. Astrophys. 388, 632-638 (2002).
51. Malavolta, L. et al. The Kepler-19 System: A Thick-envelope Super-Earth with Two Neptune-
mass Companions Characterized Using Radial Velocities and Transit Timing Variations. Astron. J.
153, 224 (2017).
52. Boisse, I., et al. Disentangling between stellar activity and planetary transits. Astron. Astrophys.
528, A4 (2011).
53. Noyes, R. W., Hartmann, L. W., Baliunas, S. L., Duncan, D. K. & Vaughan, A. H. Rotation,
convection, and magnetic activity in lower main-sequence stars. Astrophys. J. 279, 763-777 (1984).
54. Zechmeister, M. & Kürster, M. The generalised Lomb-Scargle periodogram. A new formalism
for the floating-mean and Keplerian periodograms. Astron. Astrophys. 496, 577-584 (2009).
55. Gregory, P. C. A Bayesian Analysis of Extrasolar Planet Data for HD 73526. Astrophys. J. 631,
1198-1214 (2005).
56. Matsumura, S., Takeda, G. & Rasio, F. A. On the origins of eccentric close-in planets.
Astrophys. J. Lett. 686, L29 (2008).
57. Lainey, V. Quantification of tidal parameters from Solar System data. Celest. Mech. Dyn. Astr.
126, 145-156 (2016).
58. Chambers, J. E. A hybrid symplectic integrator that permits close encounters between massive
bodies. Mon. Not. R. Astron. Soc. 304, 793-799 (1999).
59. Eastman, J., Gaudi, B. S. & Agol, E. EXOFAST: A Fast Exoplanetary Fitting Suite in IDL.
Publ. Astron. Soc. Pac. 125, 83 (2013).
60. Haywood, R. et al. Planets and stellar activity: hide and seek in the CoRoT-7 system. Mon. Not.
R. Astron. Soc. 443, 2517-2531 (2014).
61. Feroz, F., Hobson, M. P., Cameron, E. & Pettitt, A. N. Importance Nested Sampling and the
MultiNest Algorithm. Preprint at https://arxiv.org/abs/1306.2144 (2014).
62. Buchner, J. et al. X-ray spectral modelling of the AGN obscuring region in the CDFS: Bayesian
model selection and catalogue. Astron. Astrophys. 564, A125 (2014).
63. Kass, R. E. & Raftery, A. E. Bayes Factors. J. Am. Stat. Assoc. 90, 773-795 (1995).
64. Xie, J.-W. et al. Exoplanet orbital eccentricities derived from LAMOST-Kepler analysis. P.
Natl. Acad. Sci. USA 113, 11431-11435 (2016).
65. Borsato, L. et al. TRADES: A new software to derive orbital parameters from observed transit
times and radial velocities. Revisiting Kepler-11 and Kepler-9. Astron. Astrophys. 571, A38 (2014).
66. Bonomo, A. S. et al. Improved parameters of seven Kepler giant companions characterized with
SOPHIE and HARPS-N. Astron. Astrophys. 575, A85 (2015).
67. Mandel, K. & Agol, E. Analytic light curves for planetary transit searches. Astrophys. J. 580,
L171-L175 (2002).
68. Kipping, D. Efficient, uninformative sampling of limb darkening coefficients for two-parameter
laws. Mon. Not. R. Astron. Soc. 435, 2152-2160 (2013).
69. Claret, A. Testing the limb-darkening coefficients measured from eclipsing binaries. Astron.
Astrophys. 482, 259-266 (2008).
70. Fogtmann-Schulz, A. et al. Accurate Parameters of the Oldest Known Rocky-exoplanet Hosting
System: Kepler-10 Revisited. Astrophys. J. 781, 67 (2014).
71. Lithwick, Y., Xie, J. & Wu, Y. Extracting Planet Mass and Eccentricity from TTV Data.
Astrophys. J. 761, 122 (2012).
72. Mills, S. M. et al. A resonant chain of four transiting, sub-Neptune planets. Nature 533, 509-512
(2016).
73. Brugger, B., Mousis, O., Deleuil, M. & Deschamps, F. Constraints on super-Earth interiors from
stellar abundances. Astrophys. J. 850, 93 (2017).
74. Grossman, L. Condensation in the primitive solar nebula. Geochim. Cosmochim. Ac. 36, 597-
619 (1972).
75. Lopez, E. D. Born dry in the photoevaporation desert: Kepler's ultra-short-period planets formed
water-poor. Mon. Not. R. Astron. Soc. 472, 245-253 (2011).
76. Miguel, Y., Kaltenegger, L., Fegley, B. & Schaefer, L. Compositions of Hot Super-earth
Atmospheres: Exploring Kepler Candidates. Astrophys. J. Lett. 742, L19 (2011).
77. Kite, E. S., Fegley, B., Jr., Schaefer, L. & Gaidos, E. Atmosphere-interior Exchange on Hot,
Rocky Exoplanets. Astrophys. J. 828, 80 (2016).
78. Schlichting, H. E., Sari, R. & Yalinewich, A. Atmospheric mass loss during planet formation:
The importance of planetesimal impacts. Icarus 247, 81-94 (2015).
79. Zahnle, K. J. & Catling, D. C. The Cosmic Shoreline: The Evidence that Escape Determines
which Planets Have Atmospheres, and what this May Mean for Proxima Centauri B. Astrophys. J.
828, 80 (2017).
80. Springel, V. The Cosmological Simulation Code GADGET-2. Mon. Not. R. Astron. Soc. 364,
1105-1134 (2005).
81. Ćuk, M. & Stewart, S. T. Making the Moon from a Fast-Spinning Earth: A Giant Impact
Followed by Resonant Despinning. Science 338, 1047 (2012).
82. Valencia, D., O'Connell, R. J. & Sasselov, D., Internal Structure of Massive Terrestrial Planets.
Icarus 181, 545-555 (2006).
83. Benz, W. & Asphaug, E. Catastrophic Disruptions Revisited. Icarus 142, 5-20 (1999).
Supplementary Figure 1: Transits of the Kepler-107 planets. Phase-folded transits of the four
Kepler-107 planets with the best-fit model (red solid line) and residuals. For illustration purposes,
data were binned in phase bins of 30, 45, 85, and 110 s for planets b, c, d, and e.
Supplementary Figure 2: Kepler light curve. Kepler long-cadence photometric measurements
showing low-amplitude variations due to the rotational modulation of small photospheric active
regions.
Supplementary Figure 3: Weighted autocorrelation function of the Kepler light curve. The peak
at ~14 d might be the stellar rotation period.
Supplementary Figure 4: Peakbagging fit for Kepler-107. The power density spectrum centred on
the region of excess power from solar-like oscillations is shown in grey, with a 1μHz smoothed
version overlain in black. The best fitting model from the peakbagging is shown in red. The
frequencies and angular degree of the individual modes are indicated by the coloured markers.
Supplementary Figure 5: HARPS-N radial-velocity time series. Radial velocities gathered with
the HARPS-N high-accuracy and high-precision spectrograph at the Telescopio Nazionale Galileo
(La Palma, Spain) as a function of time.
Supplementary Figure 6: Generalized Lomb-Scargle periodograms of the radial-velocity and
activity indicator time series measured by HARPS-N. The panels show, from top to bottom, the
periodograms of the RV time series, the spectral window centred at the highest peak at 13.8 d in the
RV, the periodograms of the activity indexes log(R'HK), full-width at half maximum (FWHM) and
bisector (BIS) of the averaged line profile. Vertical dashed lines indicate the orbital periods of
Kepler-107b (green), c (red), d (pink), and e (blue). The horizontal dotted lines show the theoretical
false alarm probabilities of 0.1% and 1%.
Supplementary Figure 7: Generalized Lomb-Scargle periodograms of real and simulated
radial-velocity time series. The top panel shows the periodogram of the HARPS-N RV data (the
same as in previous figure). The other panels display, from top to bottom, the averaged
periodograms of simulated RVs at the real observation epochs by assuming three different values
for the mass of Kepler-107b (see Methods): Mp,b=Mp,c ( 9.4 M ⊕ ) , Mp,b=0.5 Mp,c ( 4.7 M ⊕ ), and
Mp,b=0.33 Mp,c ( 3.1 M ⊕ ). These periodograms are less noisy than the real one (top panel) mainly
because of the averaging effect. Vertical dashed lines indicate the orbital periods of Kepler-107b
(green), c (red), d (pink), and e (blue).
Supplementary Figure 8: Transit Timing Variations. The panels show, from top to bottom, the
variations of the mid-transit epochs of Kepler-107b, c, and e. Those of planet d could not be
computed because of the very low signal-to-noise ratio of its individual transits. At the achieved
precision, no significant trends are seen.
Supplementary Table 1: Priors on the free parameters of the two radial-velocity models.
Parameter
4 Keplerians + 1 Sinusoid 4 Keplerians + Gaussian
Process regression
Pb (d)
N(3.1800218, 2.9x10-6)
N(3.1800218, 2.9x10-6)
Tc,b - 2,450,000 (BJDTDB)
N(5701.08414, 3.7x10-4)
N(5701.08414, 3.7x10-4)
Kb (m/s)
Pc (d)
U[0, +∞]
U[0, 10]
N(4.901452, 1.0x10-5)
N(4.901452, 1.0x10-5)
Tc,c - 2,450,000 (BJDTDB)
N(5697.01829, 7.9x10-4)
N(5697.01829, 7.9x10-4)
Kc (m/s)
Pd (d)
U[0, +∞]
U[0, 10]
N(7.95839, 1.2x10-4)
N(7.95839, 1.2x10-4)
Tc,d - 2,450,000 (BJDTDB)
N(5702.9547 ± 0.0060)
N(5702.9547 ± 0.0060)
Kd (m/s)
Pe (d)
U[0, +∞]
U[0, 10]
N(14.749143, 1.9x10-5)
N(14.749143, 1.9x10-5)
Tc,e - 2,450,000 (BJDTDB)
N(5694.48550, 4.6x10-4)
N(5694.48550, 4.6x10-4)
Ke (m/s)
Psin (d)
T0,sin - 2,450,000 (BJDTDB)
Ksin (m/s)
h (m/s)
θ (d)
w (d)
λ (d)
RV jitter (m/s)
Vr (km/s)
U[0, +∞]
U(13.6, 14.6)
U[0, +∞]
U[0, +∞]
U[0, +∞]
U[-∞, +∞]
Notes:
N(μ, σ): normal distribution with mean μ and standard deviation σ
U[a, b]: uniform distribution between a and b values
U[0, 10]
U[0, 10]
U[12, 16]
U[0, 1]
U[0, 1500]
U[0, 10]
U[5500, 5750]
Supplementary Table 2: Results of the radial-velocity modelling. Error bars and upper limits
refer to 1σ uncertainties.
Parameter
4 Keplerians + 1 Sinusoid 4 Keplerians + Gaussian
Process regression
1.32 ± 0.57
3.06 ± 0.57
< 1.1
1.95 ± 0.82
3.51 ± 1.52
9.39 ± 1.77
< 3.8
8.6 ± 3.6
+0.25
14.10 −0.30
6827.7 ±1.3
2.41 ±0.70
Kb (m/s)
Kc (m/s)
Kd (m/s)
Ke (m/s)
M p , b( M ⊕)
M p , c(M ⊕)
M p , d ( M ⊕)
M p , e(M ⊕)
Psin (d)
T0,sin - 2,450,000 (BJDTDB)
Ksin (m/s)
h (m/s)
θ (d)
w (d)
λ (d)
RV jitter (m/s)
Vr (m/s)
< 0.8
5644.23 ± 0.45
1.39 ± 0.50
3.15 ± 0.50
< 0.9
2.56 ± 0.71
3.67 ± 1.32
9.64 ± 1.49
< 3.1
11.3 ± 3.1
+1.33
1.77 −0.93
+0.36
14.27 −0.60
0.61 ± 0.26
+396
867−492
< 0.9
+1.14
5644.30 −0.96
Supplementary Table 3: Stellar photospheric abundances relative to the Sun.
Element [X/H] Abundance [dex] Number of lines
NaI
MgI
AlI
SiI
CaI
ScI
ScII
TiI
TiII
MnI
CrI
CrII
VI
CoI
NiI
0.400 ± 0.070
0.358 ± 0.049
0.389 ± 0.014
0.361 ± 0.039
0.313 ± 0.031
0.441 ± 0.007
0.425 ± 0.093
0.350 ± 0.062
0.359 ± 0.051
0.437 ± 0.063
0.352 ± 0.047
0.253 ± 0.077
0.426 ± 0.010
0.452 ± 0.028
0.406 ± 0.025
3
3
2
14
12
3
6
24
6
5
21
3
6
8
40
Supplementary Table 4: Measurements of radial velocity and activity indexes.
Time (BJDUTC-
2,450,000)
6829.641319
6831.546838
6832.592840
6845.598973
6846.641739
6847.633230
6849.586407
6850.637493
6851.632235
6852.633376*
6853.635316*
6862.570003
6863.531728
6864.538823
6866.541785
7180.680532*
7181.593700
7182.557541
7183.584880
7185.572111
7186.594367
7188.656460
7189.648563
7190.663098
7191.663534
7192.660186
7193.662207
7195.654431
7221.605328
7222.541535
7223.715037
7227.604988
7228.609294
7230.660140
7254.608600
7256.614319
7257.617068
RV (m/s)
σRV (m/s)
FWHM (m/s)
BIS (m/s)
log(R'HK)
σlog(R'HK)
5640.11
5639.56
5649.14
5640.18
5637.14
5643.25
5642.12
5639.78
5646.18
5654.00
5652.50
5646.30
5647.75
5647.43
5642.30
5643.37
5650.47
5646.70
5635.22
5646.34
5640.02
5635.67
5638.90
5647.35
5642.24
5642.85
5640.42
5651.19
5642.08
5642.96
5634.17
5650.87
5648.84
5644.82
5657.08
5640.75
5644.00
3.97
3.26
3.77
3.01
3.74
6.30
5.67
5.61
3.42
5.06
4.14
4.13
4.48
3.19
5.84
3.72
5.40
3.31
4.96
4.81
3.18
4.57
7.85
3.43
3.32
3.47
3.09
4.00
2.91
3.84
9.55
3.75
5.40
5.03
5.64
5.04
6.71
7796.31
7812.65
7803.99
7820.68
7819.36
7823.76
7824.36
7806.19
7809.08
7797.71
7812.85
7816.29
7837.46
7803.21
7815.92
7799.86
7836.01
7814.09
7831.50
7799.70
7783.43
7782.68
7795.30
7799.31
7792.51
7809.78
7796.18
7815.44
7782.88
7780.88
7802.50
7791.80
7818.43
7818.98
7831.29
7799.96
7822.25
24.90
8.89
6.29
13.08
7.02
-5.93
3.06
18.41
-3.65
5.36
12.61
6.92
19.71
6.28
14.80
17.13
-10.56
5.08
30.06
-2.05
0.97
2.48
-18.51
5.55
10.11
12.46
8.44
15.26
6.56
17.27
8.70
1.63
8.34
7.37
4.92
-0.71
0.99
-4.994
-5.058
-5.046
-5.084
-5.037
-5.336
-5.085
-5.029
-5.078
-5.036
-4.955
-5.234
-5.111
-5.079
-5.266
-5.060
-5.007
-5.194
-5.104
-5.070
-5.139
-4.955
-5.248
-5.060
-5.052
-5.233
-5.051
-5.192
-5.155
-5.089
-4.841
-5.013
-5.067
-5.057
-4.927
-4.984
-5.189
0.080
0.070
0.086
0.066
0.076
0.320
0.154
0.132
0.078
0.128
0.078
0.148
0.124
0.072
0.241
0.083
0.123
0.098
0.129
0.123
0.083
0.092
0.326
0.080
0.073
0.117
0.064
0.124
0.075
0.100
0.187
0.078
0.141
0.136
0.123
0.107
0.272
Continued on next page
Supplementary Table 4 - continued from previous page
Time (BJDUTC-
2,450,000)
7267.552112
7269.515188
7270.498495
7271.501443
7272.540536
7273.516979
7301.493922
7302.495732
7498.703402
7499.715106*
7521.650229
7522.698371
7525.687954
7526.688837
7527.660506*
7528.680232*
7529.691289*
7530.694093*
7557.662814*
7558.633191*
7559.616508*
7565.585627
7566.606557
7573.530542
7573.551734
7574.521167
7574.541491
7576.512913
7576.533388
7602.413100
7602.433285
7614.424583*
7614.447303*
7617.419103*
7617.441696*
7618.415801*
7618.437109*
7619.391064*
RV (m/s)
σRV (m/s)
FWHM (m/s)
BIS (m/s)
log(R'HK)
σlog(R'HK)
5647.20
5653.99
5641.88
5641.57
5636.45
5647.84
5638.15
5631.49
5650.04
5648.31
5642.85
5646.94
5644.80
5635.56
5644.00
5647.21
5649.94
5641.97
5649.88
5641.31
5650.48
5644.46
5644.40
5647.73
5640.74
5638.21
5633.35
5645.31
5645.57
5644.28
5639.18
5641.24
5646.04
5647.99
5647.10
5641.99
5637.68
5635.79
3.78
3.99
3.70
3.12
5.08
3.59
3.60
5.92
6.04
5.01
4.35
3.91
5.77
6.95
5.12
4.63
5.25
6.03
5.14
6.84
7.19
8.22
3.11
4.91
4.42
3.49
4.56
4.33
3.77
7.23
5.34
3.79
3.67
4.95
5.20
5.57
7.67
3.50
7815.64
7817.17
7804.91
7805.10
7798.88
7818.20
7808.70
7812.62
7804.65
7771.64
7778.49
7795.96
7779.45
7799.03
7784.03
7789.27
7792.05
7822.29
7802.55
7794.55
7802.48
7799.69
7785.22
7805.09
7774.72
7803.44
7787.33
7784.56
7793.50
7787.16
7809.63
7795.82
7811.11
7785.28
7770.38
7775.99
7814.93
7804.13
13.86
-1.37
13.73
17.01
12.80
-0.50
15.82
32.53
24.01
23.71
6.61
27.16
8.90
13.43
8.00
23.25
33.71
25.68
-7.13
41.52
7.84
9.43
1.69
7.86
23.15
12.80
14.86
9.36
9.41
9.54
-3.55
7.68
-12.32
-20.50
-36.37
-64.62
-16.78
-17.10
-5.131
-5.013
-5.033
-5.309
-5.049
-5.125
-5.028
-5.164
-4.905
-5.449
-5.012
-5.089
-4.966
-4.961
-5.262
-5.118
-4.968
-5.242
-4.955
-5.019
-5.103
-5.093
-5.085
-5.075
-4.964
-5.061
-5.080
-5.014
-6.314
-5.147
-5.290
-5.100
-5.032
-4.968
-5.158
-5.148
-5.062
-5.113
0.101
0.072
0.059
0.247
0.074
0.091
0.137
0.200
0.091
0.235
0.084
0.161
0.156
0.105
0.182
0.157
0.134
0.202
0.146
0.157
0.245
0.071
0.131
0.111
0.064
0.108
0.104
0.072
3.430
0.166
0.140
0.088
0.116
0.107
0.170
0.244
0.076
0.085
Continued on next page
Supplementary Table 4 - continued from previous page
Time (BJDUTC-
2,450,000)
7619.413876*
7651.362036*
7651.382707*
7652.360994
7652.382047
7653.365426
7653.386397
7654.364440
7654.385528
7655.403210
7655.425003
7656.357283
7656.378463
7658.395272
7658.415573
7659.433207
7659.454318
7661.374576
7661.395478
7669.359450
7669.379947
7670.382276
7670.403398
7671.353842
7671.374605
7672.357559
7672.378241
7673.359389
7673.380719
7699.345433
7706.370203
7721.345665
7727.324829
7728.330556
7729.315081
7861.716662
7863.730901
7864.667902
RV (m/s)
σRV (m/s)
FWHM (m/s)
BIS (m/s)
log(R'HK)
σlog(R'HK)
5640.07
5652.12
5649.98
5649.98
5650.90
5640.99
5649.11
5644.77
5646.10
5646.98
5642.61
5650.08
5649.12
5634.23
5632.06
5636.31
5632.62
5642.01
5640.21
5645.77
5644.30
5645.13
5635.47
5647.08
5649.73
5648.41
5643.34
5643.28
5641.43
5649.22
5652.20
5641.33
5640.16
5646.56
5636.67
5650.46
5643.20
5632.46
3.47
4.11
3.65
3.59
3.40
6.26
6.68
4.48
4.99
5.81
7.34
3.71
3.92
7.73
7.95
4.74
5.13
3.19
3.19
3.27
3.44
5.63
5.21
3.87
3.38
3.44
3.53
4.03
3.91
4.34
6.33
5.95
3.99
5.57
5.75
6.27
5.73
3.98
7788.45
7811.95
7783.03
7808.77
7794.89
7803.50
7796.38
7791.02
7791.52
7786.50
7795.17
7767.62
7808.16
7794.18
7775.04
7798.11
7772.62
7790.73
7790.11
7793.04
7767.81
7807.98
7769.91
7778.44
7795.49
7760.09
7771.20
7797.50
7802.90
7750.61
7790.39
7780.78
7813.98
7793.96
7820.37
7803.22
7814.61
7807.42
-0.28
7.79
1.69
6.72
9.01
12.56
16.06
-2.01
-3.24
16.71
24.01
4.63
-30.68
14.02
-3.57
-5.26
-4.62
10.60
5.15
-0.61
-0.68
9.98
5.61
5.99
12.10
3.41
0.38
8.51
21.48
-15.18
10.95
10.85
12.55
-4.52
-0.95
-14.40
6.50
19.87
-5.117
-4.985
-5.246
-4.996
-5.092
-4.977
-5.111
-5.321
-5.124
-5.268
-5.045
-5.208
-4.924
-5.079
-5.136
-5.044
-5.186
-5.039
-5.176
-5.400
-5.190
-5.114
-4.935
-5.163
-5.121
-4.991
-5.073
-5.264
-5.052
-5.018
-5.000
-5.062
-4.976
-4.949
-5.113
-5.088
-5.075
-5.195
0.109
0.070
0.120
0.061
0.184
0.151
0.120
0.226
0.178
0.339
0.081
0.131
0.147
0.217
0.141
0.122
0.090
0.065
0.089
0.160
0.197
0.150
0.069
0.094
0.090
0.071
0.089
0.133
0.090
0.143
0.130
0.099
0.123
0.111
0.175
0.130
0.082
0.272
Continued on next page
Supplementary Table 4 - continued from previous page
Time (BJDUTC-
2,450,000)
RV (m/s)
σRV (m/s)
FWHM (m/s)
BIS (m/s)
log(R'HK)
σlog(R'HK)
7865.685104
5631.66
7.97
7856.53
6.00
-4.777
0.160
Notes:
*: observations corrected for moonlight contamination.
|
1101.1994 | 1 | 1101 | 2011-01-11T00:06:35 | Prospects of the Detection of Circumbinary Planets With Kepler and CoRoT Using the Variations of Eclipse Timing | [
"astro-ph.EP"
] | In close eclipsing binaries, measurements of the variations in binary's eclipse timing may be used to infer information about the existence of circumbinary objects. To determine the possibility of the detection of such variations with CoRoT and Kepler space telescopes, we have carried out an extensive study of the dynamics of a binary star system with a circumbinary planet, and calculated its eclipse timing variations (ETV) for different values of the mass-ratio and orbital elements of the binary and the perturbing body. Here, we present the results of our study and assess the detectability of the planet by comparing the resulting values of ETVs with the temporal sensitivity of CoRoT and Kepler. Results point to extended regions in the parameter-space where the perturbation of a planet may become large enough to create measurable variations in the eclipse timing of the secondary star. Many of these variations point to potentially detectable ETVs and the possible existence of Jovian-type planets. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2002)
Printed 30 October 2018
(MN LATEX style file v2.2)
Prospects of the Detection of Circumbinary Planets With
Kepler and CoRoT Using the Variations of Eclipse Timing
R. Schwarz 1⋆, N. Haghighipour2, S. Eggl1, E. Pilat-Lohinger1 and B. Funk3
1Institute for Astronomy, University of Vienna, A-1180 Vienna, Turkenschanzstrasse 17, Austria
2Institute for Astronomy and NASA Astrobiology Institute, University of Hawaii, 2680 Woodlawn Dr., Honolulu, HI, 96822 USA
3Department of Astronomy, Eotvos University, H-1117 Budapest, P´azm´any P´eter setany 1/A, Hungary
ABSTRACT
In close eclipsing binaries, measurements of the variations in binary's eclipse tim-
ing may be used to infer information about the existence of circumbinary objects. To
determine the possibility of the detection of such variations with CoRoT and Kepler
space telescopes, we have carried out an extensive study of the dynamics of a binary
star system with a circumbinary planet, and calculated its eclipse timing variations
(ETV) for different values of the mass-ratio and orbital elements of the binary and the
perturbing body. Here, we present the results of our study and assess the detectability
of the planet by comparing the resulting values of ETVs with the temporal sensitiv-
ity of CoRoT and Kepler. Results point to extended regions in the parameter-space
where the perturbation of a planet may become large enough to create measurable
variations in the eclipse timing of the secondary star. Many of these variations point
to potentially detectable ETVs and the possible existence of Jovian-type planets.
Key words: methods: numerical, techniques: photometric, binaries: eclipsing, plan-
etary systems, planets: detection
1
INTRODUCTION
Approximately 70 percent of the main and pre-main se-
quence stars are members of binary or multiple star sys-
tems. Observational evidence indicates that many of these
systems contain potentially planet-forming circumstellar or
circumbinary disks implying that planet formation may
be a common phenomenon in and around binary stars
(Mathieu 1994; Akeson et al. 1998; Rodriguez et al. 1998;
White et al. 1999; Silbert et al. 2000; Mathieu et al. 2000;
Trilling et al. 2007).
Many efforts have been made to detect such binary-
planetary systems. In the past two decade, even though sin-
gle stars were routinely prioritized in search for extrasolar
planets, many of these efforts were successful and resulted
in the detection of approximately 40 planet-hosting binary
star systems. The discovery of these systems have led to
the speculation that many more planets may exist in and
around binaries and prompted astronomers to explore the
possibility of the detection of these planets with different
detection methods. We refer the reader to Planets in Bi-
nary Star Systems (Haghighipour 2010) for an up to date
⋆ E-mail:[email protected]
and comprehensive review of the current state of observa-
tional and theoretical research in this area.
In general, one can distinguish three types of planetary
orbits in a binary star system (Fig. 1):
(i) S-Type, where the planet orbits one of the two stars,
(ii) P-Type, where the planet orbits the entire binary,
(iii) T-Type, where the planet orbits close to one of the
two equilibrium points L4 and L5 (Trojan planets).
(2002),
Currently, the most known planets in binary sys-
tems are in S-type orbits [for more details see e.g.,
Pilat-Lohinger & Dvorak
Pilat-Lohinger et al.
(2003), and Haghighipour et al. (2010)]. The stellar sepa-
rations of many of these binaries are larger than 100 AU
implying that the perturbation of their farther companions
on the formation and dynamical evolution of planets around
their planet-hosting stars may be negligible. However, in
the past few years, radial velocity and Astrometry surveys
have been able to identify five binary star systems with
separations of approximately 20 AU where one of the stars
is host to a Jovian-type planet. These systems, namely,
GL 86 (Queloz et al. 2000; Els et al. 2001), γ Cephei
(Hatzes et al.
2004;
Raghavan et al. 2006), HD 196885 (Correia et al. 2008),
and HD 176051 (Muterspaugh et al. 2010) present unique
2003), HD 41004 (Zucker et al.
2
Schwarz, Haghighipour, Eggl, Pilat-Lohinger & Funk
the detection sensitivities of these telescopes. These authors
also suggest that if target binary stars include presumably
less stable contact binaries, the CoRoT's and Kepler's capa-
bilities of detecting of circumbinary planets increase by four
times.
In this paper, we extend the study by Sybilsky et al.
(2010) to binaries with planets in eccentric and reso-
nant orbits. It is expected that similar to the transit
timing variation method where TTV signals are strongly
enhanced when the transiting and perturbing planets
are in resonance (Agol et al. 2005; Agol & Steffen 2007;
Steffen & Agol 2005), a resonant perturbing circumbinary
planet will also produce high ETV signals. The goal of our
study is to identify regions of the parameter-space for which
this signal is within the temporal sensitivity of CoRoT and
Kepler space telescopes.
The outline of our paper is as follows. We present our
models in section 2 and discuss their stability in section 3.
Section 4 has to do with the calculations of ETVs and the
prospects of the detection of P-type planets. Section 5 con-
cludes this study by summarizing the results and discussing
their implications.
2 MODEL
We consider a close, eclipsing binary with a giant planet in
a P-type orbit. To ensure that the effect of the variations
of the mass-ratio of the binary is included in our study, we
consider the following three models where m1 and m2 are
the masses of the primary and secondary stars, respectively,
• model 1: m1 = m2 = 0.3 Msun,
• model 2: m1 = 1, m2 = 0.5 Msun,
• model 3: m1 = m2 = 1 Msun.
As most of the stars in the solar neighborhood are of spectral
type M, this choice of models ensures that low-mass binary
stars are also included in our simulations.
Eclipsing binaries are morphologically classified as
(i) detached systems, if neither component fills its
Roch lobe (separated stars),
(ii) semi-detached systems, if only one component fills
its Roche lobe, and
(iii) over contact systems, if both components exceed
their Roche lobes.
The Roche lobe marks the volume limit at which the star
may begin to lose substantial amount of matter to its com-
panion. In this study we consider a detached binary with
an initial separation of 0.05 AU. For the binary models 1 and
3, this separation corresponds to a period of approximately
5 and 3 days, respectively. As shown by Goldmann (2010),
most candidate eclipsing binaries in the discovery space of
CoRoT have periods between 1 and 10 days.
When dealing with close binaries such as the models
considered here, the intrinsic eccentricity of the binary (ebin)
can be neglected. This is due to the high probability of cir-
cularization that is caused by interstellar tidal forces. In this
study, we consider the initial orbital eccentricity of the bi-
nary to be zero or have very low values. The latter is due
to the fact that the timescale of circularization is dependent
upon specific stellar parameters (Zahn & Bouchet 1989).
Figure 1. Schematic illustration of S-type and P-type orbits in a
binary star system. The yellow and red circles represent the stars
of the binary revolving around their common center of mass (the
blue circle).
cases for the study of planet formation and dynamics in bi-
naries as the perturbation of their secondary stars will have
significant effects on the dynamics of their circumprimary
disks and their capability in forming planets.
The success of radial velocity and astrometry in detect-
ing planets in S-type orbits raises the question that whether
other techniques can also detect planets in and around bi-
nary stars. A recent success has come in the form of the
detection of the first P-type planets using the ETV method.
Modeling the variations in the eclipse timing of the binary
NN Ser, Beuermann et al. (2010) have shown that this sys-
tem is host to two planets with minimum masses of 2.2 and
6.9 Jupiter-masses in a 2:1 mean-motion resonance in cir-
cumbinary orbits.
Our work has been motivated by this discovery. NN
Ser (ab) is an eclipsing short period binary which shows
long-term ETVs. The detection of P-type planets around
this system suggests that close, eclipsing binaries may be
promising candidates in the search for new exoplanets. As
many of these binaries lie in the discovery space of CoRoT
(Goldmann 2010) and Kepler (Coughlin et al. 2010) space
telescopes, we focus our study on exploring the prospects of
the detection of such objects.
The idea of photometric detection of extrasolar plan-
ets around eclipsing binaries was first presented by
Schneider & Chevreton (1990). This idea that was later
developed by many authors
(Schneider & Doyle 1995;
Doyle et al. 1998; Doyle & Deeg 2004; Deeg et al. 2008;
Muterspaugh et al. 2010), is based on the fact that a cir-
cumbinary planet can perturb the orbit of the two stars and
create variations in their eclipse timing. The measurements
of these variations, when compared with theoretical models,
can reveal information about the mass and orbital elements
of the perturbing planet.
Recently Sybilsky et al. (2010) studied the potential of
the ETV method in detecting giant planets in circular, cir-
cumbinary orbits. These authors have shown that ground-
based photometry may have advantage over CoRoT and
Kepler space telescopes in detecting circumbinary planets.
They suggest that the selective nature of the target stars in
the fields of view of CoRoT and Kepler causes the ETV sig-
nals of eclipsing binaries in these regions to be smaller than
Circumbinary Planets and Eclipse Timing Variations
3
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
0.1
0.14
0.18
0.22
0.26
0.30
0.1
0.14
0.18
0.22
0.26
0.30
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
Semimajor Axis (AU)
Semimajor Axis (AU)
0.30
0.25
y
t
i
c
i
r
t
n
e
c
c
E
0.20
0.15
0.10
0.05
0
0.1
0.14
0.18
0.22
0.26
0.30
0.1
0.14
0.18
0.22
0.26
0.1
0.30
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
Semimajor Axis (AU)
Semimajor Axis (AU)
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
0.1
0.14
0.18
0.22
0.26
Semimajor Axis (AU)
0.1
0.30
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
Figure 2. The emax maps of a circumbinary planet in the binary model 1 (left column), 2 (middle column), and 3 (right column). The
top row shows the maps for a 1 MJ planet, the middle row is for a 5 MJ planet, and the bottom row corresponds to a 10 MJ object.
The violet and blue regions represent small values of emax and correspond to bound orbits whereas the yellow region points to escape.
3 STABILITY ANALYSIS
Prior to calculating the variations in the eclipse timing, it is
necessary to identify regions where a P-type planet can have
a stable orbit. For this purpose, we integrated the Newtonian
three-body system of the binary and its planet for different
values of the planet's mass and orbital elements. The mass
of the planet was taken to be 1, 5, and 10 Jupiter-masses.
The values of its semimajor axis and orbital eccentricity with
respect to the barycenter of the system were varied between
a = 0.1 − 0.3 AU and e = 0 − 0.3 in increments of ∆a = 0.01
AU and ∆e = 0.01, respectively.
Numerical integrations were carried out using the Lie-
series method and Bulirsch-Stoer integrator (Lichtenegger
1984; Hanslmeier & Dvorak 1985; Eggl & Dvorak 2010). We
integrated the system for 104 years which corresponds to the
following number of periods of the planet at 0.1 AU:
• model 1: 2.45 × 105 periods,
• model 2: 3.90 × 105 periods,
• model 3: 4.50 × 105 periods.
The stability of the planetary motion was controlled
by measuring the value of its maximum orbital eccentricity
(emax). We monitored the changes in the eccentricity of the
planet throughout the integration and determined its high-
est value. If the orbit of the planet became parabolic (e = 1),
we considered the system to be unstable.
Figure 2 shows the values of planet's emax at different
distances from the barycenter of the binary. The left column
corresponds to the binary model 1, the middle column to
binary model 2, and the right column to binary model 3.
Also, from top to bottom, each row corresponds to the values
of emax for a 1 MJ , 5 MJ , and 10 MJ , respectively. In each
panel, the axes represent the initial values of the semimajor
axis and eccentricity of the planet at the beginning of the
integration. The color at each point depicts the maximum
value that the eccentricity of the planet acquired during the
integration. As indicated by the scale on the right side of the
figure, yellow corresponds to parabolic orbits and denotes
instability.
An inspection of the emax maps shown in Fig. 2 indi-
cates that for planets in circular or low eccentricity orbits,
the range of the semimajor axis for which the planetary or-
bit may be bound is large and does not change much for
different values of the planet's mass. For instance, as shown
by the top left panel of this figure, the boundary of the
unstable zone for a 1 MJ planet in a circular orbit in the
binary model 1 is interior to 0.1 AU. This suggests that all
circular orbits with semimajor axes larger than 0.1 AU in
this model may be bound. As the mass of the planet grows
4
Schwarz, Haghighipour, Eggl, Pilat-Lohinger & Funk
(middle and bottom panels of the left column in Fig. 2),
the outer boundary of the unstable region slowly progresses
toward larger semimajor axes. Such a trend is also seen in
the emax maps of binary models 2 and 3. As shown by the
lower panels in the middle and right columns, the bound-
ary of the unstable zone for circular orbits is slightly shifted
outward to 0.11 AU. These results are consistent with the
results of the stability analysis of a test particle in a cir-
cumbinary orbit as presented by Dvorak et al. (1989) and
Holman & Wiegert (1999). According to the estimate of the
boundary of the stable zone as given by these authors, the
critical distance beyond which the orbit of a Jovian planet
in all our three binary models will be stable is in the range
of 0.1 to 0.12 AU.
While the value of emax can be used to identify parabolic
(unstable) orbits, it cannot be used as a rigorous indica-
tor of planet's orbital stability. A low value of emax im-
plies a bound planetary orbit. But that is only for the du-
ration of the integration, and there is no guarantee that
the orbit of the planet will stay bound for a long time. In
order to determine the orbital stability of the planet, we
used the Fast Lyapunov Indicator (FLI) (Froeschl´e et al.
1997). The FLI is a chaos indicator that measures the
exponential divergence of nearby trajectories and distin-
guishes between regular and chaotic motion. For details
of this technique, we refer the reader to Froeschl´e et al.
(1997), Lega & Froeschl´e (2001) and Fouchard et al. (2002).
Applications of FLI to planetary motion in binary star
systems can be found in Pilat-Lohinger & Dvorak (2002),
Pilat-Lohinger et al. (2003) and Haghighipour et al. (2010).
Figure 3 shows a sample of an FLI stability map for a 1
MJ planet in binary model 1. The colors in this figure rep-
resent the values of FLI (the logarithmic scale on the right)
which depict the degree of chaos. Dark colors correspond
to less chaotic and regular orbits. As shown by this figure,
there is a region in the parameter-space where the orbit of
the planet is likely stable (dark area between 0.1 and 0.3
AU).
A comparison between the dark region of Fig. 3 and
its corresponding region in the system's emax map in Fig. 2
(upper left panel) points to an interesting observation: even
though the maximum values of the planet's orbital eccentric-
ity can reach to high values, the orbit of the planet may still
be stable. In other words, elevated planetary eccentricities
do not necessarily correspond to instability. For instance, as
shown by the upper left panel of Fig. 2, orbits with initial
eccentricities of 0.3 may have emax values up to 0.7. Never-
theless, according to the FLI results shown in Fig. 3, they
are probably long-term stable.
Similar comparison can also be made between the
chaotic region of Fig. 3 (indicated by red and yellow col-
ors) and the system's emax map in Fig. 2. The results point
to an inverse phenomenon, that is, instability among seem-
ingly bound orbits. For instance, while for semimajor axes
ranging from a = 0.11 AU to a = 0.14 AU, and the ec-
centricities between e = 0.05 and e = 0.15, the emax map
suggests bound planetary orbits, the FLI stability map indi-
cates chaotic motion. The different color shades for chaotic
orbits in the FLI map have resulted from the fact that the
degree of divergence depends on the initial conditions of the
orbits. Even though this shading may suggest less chaotic
'islands' in parameter-space, when compared to the corre-
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
Figure 3. The FLI stability map for a 1 MJ P-type planet in the
binary model 1. The violet and blue regions (small values of the
FLI) correspond to regions of regular motion, and red and yellow
denote chaotic orbits.
sponding emax map, one can see that orbits inside the chaotic
border zone become parabolic.
Using the combination of emax and FLI analyses, we
identified the stable planetary orbits in all our three binary
models. As an example, the number of stable circular or-
bits are shown in Table 1. As expected, stability is almost
independent of the mass-ratio of the binary.
We also carried out simulations for non-zero values of
the binary's eccentricity. Figure 4 shows the maps of emax for
a 1 MJ planet in the binary model 1. The binary eccentricity
was chosen to be ebin = 0.05, 0.1, and 0.15. Table 2 shows
the number of stable orbits for these simulations. As shown
here, the unstable region expands out by more than two folds
when ebin = 0.1. Our analysis indicated that for the value
of the binary eccentricity ebin = 0.2, only 4 orbits remained
stable (Table 2).
The results of our stability analysis allow us to focus
our calculations of ETVs on the region of the parameter-
space where P-type planets have stable orbits. As shown,
the planet's stable region shrinks for large values of its ec-
centricity as well as the eccentricity of the binary. Initial
binary eccentricities beyond ebin = 0.15 move this region
to large distances (a = 0.24 AU) where the perturbing ef-
fect of the planet on the dynamics of the binary becomes
negligible (Fig. 4). The closest possible stable orbit for a
planet (a = 0.1 AU) is when the planet is Jupiter-mass, and
both the planet and binary have circular orbits. As a result,
for the purpose of calculating ETVs, we only consider fully
circularized binaries.
4 CALCULATION OF ECLIPSE TIMING
VARIATIONS
As indicated by the results of the stability analysis, a Jovian
planet in a P-type orbit may be stable in the vicinity of a
binary star system. Although small, the gravitational per-
turbation of this planet may affect the motions of the two
stars and cause their orbits to deviate from Keplerian. In an
eclipsing binary, these deviations result in variations in the
time and duration of the eclipse. Similar to the variations in
the transit timing of a planet due to a second perturber
(Miralda-Escud´e & Adams 2005; Holman & Murray 2005;
Agol & Steffen 2007; Kipping 2009a,b; Ford & Holman
Circumbinary Planets and Eclipse Timing Variations
5
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
y
t
i
c
i
r
t
n
e
c
c
E
0.30
0.25
0.20
0.15
0.10
0.05
0
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.1
0.14
0.18
0.22
0.26
0.30
Semimajor Axis (AU)
Figure 4. The emax maps of a 1 MJ planet in the binary model
1 with ebin = 0.05 (top), 0.1 (middle), and 0.15 (bottom).
Table 1. Number of stable orbits of all three binary models for
ebin = 0.
Model
Mplanet
(MJ )
model 1
model 2
model 3
1
5
10
1
5
10
1
5
10
stable orbits
stable orbits
(number)
540
530
527
531
529
526
539
538
533
(%)
82.9
81.4
80.9
81.5
81.2
80.7
82.6
82.6
81.8
Table 2. Number of stable orbits for binary model 1 and for
different values of ebin.
ebin Mplanet
(MJ )
stable orbits
(numbers)
0.00
0.05
0.10
0.15
0.20
1
1
1
1
1
540
407
270
127
4
stable orbits
(%)
82.9
62.5
41.4
19.5
0.6
2007), the variations in eclipse timing can be used to in-
fer information about the mass and orbital elements of the
circumbinary planet. In this section, we calculate the eclipse
timing variations of our model binary star systems for differ-
ent values of the mass and orbital parameters of the binary
and its P-type planet. Our goal is to identify a range of
these parameters for which the magnitude of ETVs will be
within the temporal sensitivity of CoRoT and Kepler space
telescopes.
We simulated the dynamics of our binary models and
their P-type planets in a barycentric coordinate system.
Simulations were carried out for 1 year corresponding to ap-
proximately 70 transits of the secondary star. We calculated
ETVs by determining the difference between the eclipse tim-
ing (t1) of the unperturbed system (star-star) and its corre-
sponding value (t2) in the perturbed case (star-star-planet)1.
Additionally, we subtracted the planetary induced constant
rate of apsidal precession. We also Fourier-analyzed the re-
sulting signal and compared the superposition of the main
frequencies to simple estimates on maximum amplitudes
[(dtmax − dtmin)/2] for the entire duration of integration.
In cases where the differences between Fourier-composition-
derived ETVs and maximum amplitudes were lower than
10%, we chose to present the ETVs' maximum amplitudes.
4.1 Implication for the detection of circumbinary
planets
As mentioned earlier, we would like to identify regions of the
parameter-space for which the ETV signals of an eclipsing
binary will be detectable by CoRoT and Kepler space tele-
scopes. Since as indicated by the stability analysis, stable
planetary orbits move to large distances as the eccentricity
of the binary increases, we limited our calculations to circu-
lar systems. We recall that in all our models, the separation
of the binary is 0.05 AU.
For the planet's motion, we considered both resonant
and non-resonant orbits. Similar to transit timing variation,
ETV signals are strongly enhanced when the P-type planet
and the binary are in a mean-motion resonance (MMR).
Results of our stability analysis indicate that the location of
the 2:1 MMR at 0.079 AU is too close to the binary's center
of mass to be stable (Dvorak et al. 1989, Holman & Wiegert
1999). We, therefore, considered cases where the planet is in
a 3:1 (0.104 AU) and a 4:1 (0.126 AU) MMR.
1 We chose to integrate the binary two body problem instead of
taking analytical solutions in order to minimize the influence of
the numerical integration algorithm on the results.
6
Schwarz, Haghighipour, Eggl, Pilat-Lohinger & Funk
For the non-resonant orbits, we chose the semimajor
axis of the planet to have the values of 0.2 AU and 0.3
AU. The eccentricity of the planet's orbit in all cases was
chosen to be 0 or 0.15, and its angular variables (inclination,
mean-anomaly, longitude of the ascending node, argument
of pericenter) were set to zero.
Table 3 shows the results of the calculation of ETVs for
a circular binary. Even though MMRs did not have too much
influence on the stability of the system, the most prominent
ETVs were found at 3:1 and 4:1 resonances. Figure 5 shows
these ETVs for a 1 Jupiter-mass planet.
In order to assess the detectability of our ETV signals
with CoRoT and Kepler, we compared our results with the
values of the Detectable Timing Amplitudes (DTA) of these
telescopes. As shown by Sybilsky et al. (2010), given the
typical photometric error, the value of the DTA for CoRoT
is approximately 4 sec for a 12 mag star and 16 sec for
a star with 15.5 mag. Similar analysis suggests that stars
with magnitudes between 9 and 14.5 mag will have DTAs
ranging from 0.5 to 4 sec with Kepler.
Figure 6 shows a comparison between the values of DTA
calculated by Sybilsky et al. (2010) and the amplitudes of
ETVs obtained from our simulations. The x-axes in this fig-
ure represent the planet's initial semimajor axis and the y-
axes show the maximum amplitude of ETVs. In cases where
the value of ETV is larger than 40 sec, this value is indicated
on the top of its corresponding bar. As a point of compar-
ison, the maximum values of the DTAs for CoRoT (16 sec
for a 15.5 mag star) and Kepler (4 sec for a 14.5 mag star)
are also shown. When the ETV signal is higher than these
observational thresholds, we assume that a planet may be
indirectly detectable via eclipse timing of the secondary star.
As shown in this figure, planets in mean-motion res-
onances present the highest probability of detection. This
is expected as these planets create the largest variations in
eclipse timing. Also, as indicated by the left column, binary
model 1 with low-mass stars present the best prospect for
detecting P-type Jovian planets. As the masses of the binary
stars increase, the prospect of detection shifts toward larger
planets. For instance, in binary model 3 (right column), a 5
Jupiter-mass planet can be detected in orbits smaller than
0.2 AU. As expected, the highest prospect of detection is
for a 10 Jupiter-mass object (lower right panel). Figure 6
also shows that while planets in circular orbits produce the
largest ETVs when in resonance, slight eccentricity in the
orbit of the planet increases the prospect of its detection
and extends its detectability to large distances. The latter
is more pronounced for larger planets around more massive
binaries as in these cases, the planetary orbit stays stable
when its orbital eccentricity is slightly increased.
Table 3. Systems with ETVs that are potentially detectable with
CoRoT and Kepler.
Model
Mplanet
(MJ )
model 1
model 2
model 3
1
1
5
5
5
10
10
10
1
5
10
10
1
5
10
10
a
e
(AU)
0.104
0.126
0.126
0.2
0.3
0.126
0.2
0.3
0.126
0.126
0.126
0.2
0.104
0.126
0.126
0.2
0
0, 0.15
0, 0.15
0, 0.15
0.15
0,0.15
0, 0.15
0.15
0.15
0, 0.15
0
0, 0.15
0
0, 0.15
0, 0.15
0.15
ETV
(sec)
316
40, 69
179, 349
25, 51
28
339, 730
46, 97
54
32
55, 161
108
17, 21
51
31, 59
60, 117
16
Figure 5. The graphs of ETVs for a 1 MJ planet in a circular
orbit in the binary model 3. The planet is in a 3:1 MMR in the
top panel and a 4:1 MMR in the bottom one.
5 DISCUSSION
We have studied the prospects of the detection of circumbi-
nary planets with CoRoT and Kepler space telescopes using
the variations of binary eclipse timing. The uninterrupted
high precision photometry of these telescopes (Koch et al.
2004; Alonso et al. 2008) has given them unique capabil-
ity for detecting small variations in eclipse timing measure-
ments. We have calculated such variations in several bi-
nary models with circumbinary planets and compared the
results with the temporal sensitivity of these telescopes.
As expected, the prospect of detection is higher for plan-
ets in resonant orbits as these objects create larger ETVs.
This is more pronounced when the stars of the binary have
low masses. This result is consistent with the findings by
Schneider & Doyle (1995). However, it is important to note
that low-mass binaries may be hard to identify.
Our study indicates that around solar-mass binaries,
Circumbinary Planets and Eclipse Timing Variations
7
Figure 6. Graphs of the maximum amplitude of ETVs for all three binary models with ebin = 0. The locations of bars on the horizontal
axis correspond to 3:1 MMR at 0.104 AU, 4:1 MMR at 0.126 AU, and two non-resonant orbits at 0.2 and 0.3 AU. Graphs are shown
for two values of the planet's orbital eccentricity (0 and 0.15). As a point of comparison, the maximum values of DTAs for CoRoT and
Kepler (Sybilsky et al. 2010) are also shown.
planets with sizes down to 1 Jupiter-mass can produce de-
tectable signals. In this case, planets have to move in almost
circular orbit close (a ∼ 2 abin) to binary's center of mass in
order to maintain stability. More massive planets on higher
eccentricities may orbit the binary at distances up to ap-
proximately 6 abin and still be able to produce ETVs that
are detectable by Kepler.
Result of our simulations indicate that although the or-
bital stability of the planet is strongly affected by increasing
its eccentricity and the eccentricity of the binary, slight de-
viations from circular orbits, in particular in the orbit of
the planet, result in its periodic close approaches to the bi-
nary and creating large ETVs. Our study suggests that when
not in a resonance, in general, giant planets on slightly ec-
centric circumbinary orbits show bigger prospects for hav-
ing detectable ETVs. With their high precision photometry
and long duration of operation, CoRoT and Kepler are well
suited for indirect planet detection via ETVs, and have the
capability of detecting such planets within the durations of
their operation.
ACKNOWLEDGMENTS
RS acknowledges supports through the MOEL grant of the
OFG (Project MOEL 386) and the FWF project P18930.
NH acknowledges support from the NASA Astrobiology In-
stitute under Cooperative Agreement NNA04CC08A at the
Institute for Astronomy, University of Hawaii, and NASA
EXOB grant NNX09AN05G. Supports are also acknowl-
edged through the Austrian FWF Erwin Schrodinger grant
no. J2892-N16 for BF, the Austrian FWF project no. P20216
for SE, and the Austrian FWF project no. P22603, P20216
for E. P-L. We thank J. Schneider and G. Wuchterl for fruit-
ful discussions.
REFERENCES
Agol, E., Steffen, J., Sari, Re´em, Clarkson, W. 2005, MN-
RAS, 359, 567
Agol, E. & Steffen J.H., 2007, MNRAS, 374, 941.
Akeson, R. L., Koerner, D. W. & Jensen, E. L. N., 1998,
ApJ, 505, 358
Alonso, R., et al. 2008, A& A, 482, L21
8
Schwarz, Haghighipour, Eggl, Pilat-Lohinger & Funk
Beuermann, K. Hessmann, F. V., Dreizler, S., Marsh, T.
R., Parsons, S. G., Winget, D. E., Miller, G. F., Schreiber,
M. R., Kley, W., Dhillon, V. S., Littlefair, S. P., Copper-
wheat, C. M., & Hermes, J. J., A&A, 521, L60
Correia, A. C. M., Udry, S., Mayor, M., Eggenberger, A.,
Naef, D., Beuzit, J.-L., Perrier, C., Queloz, D., Sivan, J.-
P., Pepe, F., Santos, N. C., & S´egransan, D. 2008, A& A,
479, 271
Coughlin, J. L., Lopez-Morales, M., Harrison, T. E., Ule,
N., Hoffman, D. I., 2010, submitted (arXiv:1007.4295v2)
Deeg, H. J., Ocana, B., Kozhevnikov, V. P., Charbonneau,
D., O ´Donovan, F. T., Doyle, L. R. 2008,A& A, 480, 563
Doyle L. R., Deeg H. J., Jenkins J. M., Schneider, J.,
Ninkov, Z., et al. 1998, Detectability of Jupiter-to-Brown-
Dwarf-Mass Companions around Small Eclipsing Binary
System, in Rebolo R., Martin, E. L., Zapatero-Osorio
M.R. (eds.) Brown Dwarfs & Extrasolar Planets. ASP
Conference Proc. Vol. 134, 224
Doyle, L. R., & Deeg, H.-J. 2004, in Norris,R., Stootman,
F., eds. Proceedings of IAU Symposium 213, Bioastron-
omy 2002: Life Among the Stars, San Francisco: Astro-
nomical Society of the Pacific, 2003., p.80
Dvorak, R., Froeschl´e, Ch.,Froeschl´e, C., 1989, A&A, 226,
335
Eggl, S., & Dvorak, R., 2010, LNP, 790, 431
Els, S. G., Sterzik, M. F., Marchis, F., Pantin, E., Endl,
Hartkopf, W. I., Boss, A. P., & Williamson, M., 2010,
MNRAS, 140, 1657
Pilat-Lohinger, E. & Dvorak,R.,2002, CeMDA, 82, 143.
Pilat-Lohinger, E., Funk, B., Dvorak, R., 2003, A&A, 400,
1085.
Queloz, D., Mayor, M., Weber, L., Bl´echa, A., Burnet, M.,
Confino, B., Naef, D., Pepe, F., Santos, N., & Udry, S.,
2000, A& A, 354, 99
Raghavan, D., Henry, T.J., Mason, B. D., Subasavage, J.P.,
Jao, W-C., Beaulieu, T. D., & Hambly, N. C., 2006, ApJ,
646, 523
Rodriguez, L. F., D'Alessio, P., Wilner, D. J., Ho, P. T. P.,
Torrelles, J. M., Curiel, S., Gomez, Y., Lizano, S., Pedlar,
A., Canto, J. & Raga, A.C.,1998, Nature, 395, 355
Schneider, J. & Chevreton, M. 1990, A & A, 232, 251.
Schneider, J. & Doyle L.R., 1995, EM&P, 71, 153.
Silbert, J., Gledhill, T., Duchene, G. & M'enard, F., 2000,
ApJL, 536, L89
Steffen, J. H., & Agol, E. 2005, MNRAS, 364, L96
Sybilsky, P., Konacki and Koz lowski, S., 2010, MNRAS,
405, 657.
Trilling, D. E., Stansberry, J. A., Stapelfeldt, K. R., Rieke,
G. H., Su, K. Y. L., Gray, R. O., Corbally, C. J., Bryden,
G., Chen, C. H., Boden, A. & Beichman, C. A., ApJ, 658,
1264
White R. J., Ghez A. M., Reid I. N. & Schultz G., 1999,
M., & Kurster, M. 2001, A& A, 370, L1
ApJ, 520, 811
Zucker, S., Mazeh, T., Santos, N.C., Udry, S., Mayor, M.,
2004, A & A, 426, 695.
Zahn, J.-P. & Bouchet, L. 1989, A& A, 223, 112
Ford, E.B., & Holman, M.J., 2007, ApJ, 664, L51
Fouchard, M., Lega E., Froeschl´e Ch., Froeschl´e C.,2002,
CeMDA, 83, 205
Froeschl´e, C., Lega E., Gonzi, R.,1997, CeMDA, 67, 41.
Goldmann, C., Secular resonances in planetary systems,
Master Thesis, University of Vienna.
Haghighipour, N., 2010, Planet in Binary Star Systems
(Berlin: Springer), ISBN: 978-90-481-8686-0
Haghighipour, N., Dvorak, R., & Pilat-Lohinger, E., 2010,
in "Planets in Binary Star Systems", N. Haghighipour
(ed.), Astrophysics and Space Science Library, Vol. 366.
Berlin: Springer, ISBN: 978-90-481-8686-0, p.285-327
Hanslmeier, A., &, Dvorak, R., 1984, A&A, 132, 203
Hatzes, A. P., Cochran, W. D., Endl, M., McArthur, B.,
Paulson, D.B., Walker, G. A. H., Campbell, B. & Yang,
S., 2003, ApJ, 599, 1383
Holman, M. J. & Murray, N. W., 2005, Science, 307, 1288
Holman, M. J. & Wiegert, P. A., 1999, AJ, 117,621.
Kipping, D.M., 2009a, MNRAS, 392, 181
Kipping, D. M., 2009b, MNRAS, 396, 1797
Koch, D. G., et al. 2004, Proc. SPIE, 5487, 1491
Lega E. & Froeschl´e, C.,2001, CeMDA, 81, 129
Lichtenegger, H. 1984, CeMDA, 34, 357.
Mathieu, R.D., 1994, ARA& A, 32, 465
Mathieu, R. D., Ghez, A. M., Jensen, E. L. N. & Simon, M.,
2000, in Protostars and Planets IV, ed. V. Mannings, A.
P. Boss, and S.S. Russell (Tucson: Univ. Arizona Press),
pp.703
Miralda-Escud´e, J. & Adams, F.C. 2005, Icarus, 178, 517
Muterspaugh, M. W., Konacki, M., Lane, B. F., & Pfahl,
E. 2010, in "Planets in Binary Star Systems", N. Haghigh-
ipour (ed.), Astrophysics and Space Science Library, Vol.
366. Berlin: Springer, ISBN: 978-90-481-8686-0, p.77-103
Muterspaugh, M. W., Lane, B. F., Kulkarni, S. R.,
Konacki, M., Burke, B. F., Colavita, M. M., Shao, M.,
|
1908.10395 | 1 | 1908 | 2019-08-27T18:26:24 | 1:1 orbital resonance of circumbinary planets | [
"astro-ph.EP"
] | The recent detection of the third planet in Kepler-47 has shown that binary stars can host several planets in circumbinary orbits. To understand the evolution of such systems we have performed two-dimensional hydrodynamic simulations of the circumbinary disc with two embedded planets for several Kepler systems. In two cases, Kepler-47 and -413, the planets are captured in a 1:1 mean-motion resonance at the planet parking position near the inner edge of the disc. The orbits are fully aligned, have mean eccentricities of about 0.25 to 0.30, and the planets are entangled in a horseshoe type of motion. Subsequent n-body simulations without the disc show that the configurations are stable. Our results point to the existence of a new class of stable resonant orbits around binary stars. It remains to be seen if such orbits exist in reality. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. 1to1resonances
August 29, 2019
©ESO 2019
1:1 orbital resonance of circumbinary planets
Anna B. T. Penzlin, Sareh Ataiee and Wilhelm Kley
Institut für Astronomie und Astrophysik, Universität Tübingen, Auf der Morgenstelle 10, D-72076 Tübingen, Germany
e-mail: {anna.penzlin, sareh.ataiee, wilhelm.kley}@uni-tuebingen.de
ABSTRACT
The recent detection of the third planet in Kepler-47 has shown that binary stars can host several planets in circumbinary
orbits. To understand the evolution of such systems we have performed two-dimensional hydrodynamic simulations of
the circumbinary disc with two embedded planets for several Kepler systems. In two cases, Kepler-47 and -413, the
planets are captured in a 1:1 mean-motion resonance at the planet parking position near the inner edge of the disc. The
orbits are fully aligned, have mean eccentricities of about 0.25 to 0.30, and the planets are entangled in a horseshoe
type of motion. Subsequent n-body simulations without the disc show that the configurations are stable. Our results
point to the existence of a new class of stable resonant orbits around binary stars. It remains to be seen if such orbits
exist in reality.
Key words. Hydrodynamics -- Binaries: general -- Planets and satellites: resonance --
1. Introduction
By now about a dozen planets orbiting around binary star
systems have been detected by the Kepler space mission. All
of these are single-planet systems, except Kepler-47 whose
third planet was recently detected (Orosz et al. 2019),
demonstrating that circumbinary systems can host multi-
ple planets. This raises the question about their formation,
evolution and stability.
Following previous approaches we assume that the plan-
ets formed further away from the central binary in more
quiet disc regions and then migrated inwards where they
are parked at the cavity edge. Using hydrodynamical disc
simulations, we modelled the migration of multiple planets
in 6 of the known Kepler circumbinary systems and discov-
ered stable coorbital resonant configurations in Kepler-47
and -413.
Thus far, 1:1 resonances have not yet been observed in
exoplanet systems but only for minor Solar System objects.
Well-knowns are the population of Trojans of Jupiter and
Saturn, and the moons Janus and Epimetheus that orbit
Saturn, which have similar mass and are on mutual horse-
shoe orbits (Yoder et al. 1983). For the latter, Rodríguez
et al. (2019) have recently presented a formation mechanism
involving migration in a circumplanetary disc. Relating to
exoplanets, formation of 1:1 resonance objects can proceed
for example in-situ (Beaugé et al. 2007), or through a con-
vergent migration process (Cresswell & Nelson 2006). In
the first scenario the planets begin on tadpole orbits, in the
latter case they start out with horseshoe motion that later
turn into tadpole as well (Cresswell & Nelson 2006).
However, continued migration of the resonant pair in
a gas disc will lead to an increase in the libration ampli-
tude which can destabilise the system (Cresswell & Nelson
2009), in particular for more massive planets (Pierens &
Raymond 2014). Stabilisation of a system can be achieved
if the pair is in resonance with a third planet (Cresswell &
Nelson 2009; Leleu et al. 2019a), whose presence may also
assist in forming the 1:1 resonance in the first place, by
providing additional perturbations. Around a binary star,
planets reach a stable parking position near the disc's inner
cavity as determined by the binary and the disc parameters.
In case of a successful capture near this position one might
expect long-term stability for such coorbital configurations
as they do not migrate any further.
For the pure n-body case, the dynamics and stability of
eccentric coplanar coorbitals around single stars has been
studied for the equal planet mass case in Leleu et al. (2018).
They show that for low mass planets around 10−5 of the
stellar mass, stable horseshoe and tadpole orbits exist up
to an orbital eccentricity of about 0.5. The stability of 1:1
resonances between low mass planets has been studied in
the special case of an equal mass binary star on a circular
orbit by Michalodimitrakis & Grigorelis (1989), who found
stability only in the zero mass limit of the planets.
Here, we present the disc driven migration of a system
of small and equal mass planets in the Kepler-47 and -413
systems. We show that in both cases stable 1:1 resonant
configurations are possible. In sect. 2 we describe our model
setup, in sect. 3 we focus on the resulting 1:1 resonances,
and in sect. 4 we discuss out results.
2. Models
To model discs around binary stars, we perform viscous
hydrodynamic simulations of thin discs in two dimensions
(2D) using a modified version of the PLUTO-code (Mignone
et al. 2007) that allows for computation on graphics pro-
cessing units (Thun et al. 2017). We use a locally isothermal
disc because the final equilibrium structure of circumbinary
discs with full radiative cooling and viscous heating is simi-
lar to a locally isothermal model using the equilibrium tem-
Article number, page 1 of 5
9
1
0
2
g
u
A
7
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
9
3
0
1
.
8
0
9
1
:
v
i
X
r
a
Kepler
MA [M(cid:12)]
MB [M(cid:12)]
abin [au]
ebin
y
r
a
n
B
i
y acav [abin]
t
i
v
a
C
ecav
Tprec [Tbin]
A&A proofs: manuscript no. 1to1resonances
413
0.82
0.54
0.10
0.04
5.18
0.45
2271
47
1.04
0.36
0.08
0.02
5.24
0.43
2573
Table 1: Parameters for Kepler-413 and -47, as used in the
simulations. Top: MA and MB denote the stellar masses
of the binary, with semi-major axis, abin, and eccentricity,
ebin. Bottom: Mean equilibrium orbit parameters of the
inner cavity of the disc without planets, as obtained by the
hydrodynamical evolution. Tprec is the precession period,
acav the cavity semi-major axis, and ecav its eccentricity.
Fig. 2: Density distribution at the time right before the
capture (left) and in the coorbital configuration (right) for
Kepler-47 (top) and -413 (bottom). The white ellipse indi-
cates the location of the cavity in the disc and the light blue
ellipses show the orbits of both planets, where the yellow
dots mark their positions. The orbital parameters of the
outer planet are quoted at the top left corner.
of the cavities, calculated by fitting the inner disc edge (see
Thun et al. (2017) for details), are given in the lower block
of Table 1. The cavity size and eccentricity is determined by
the binary eccentricity and mass ratio, the disc height and
viscosity. In agreement with Thun & Kley (2018), systems
with small ebin show very large cavity eccentricities. In our
case we obtain ecav ∼ 0.45, see Table 1.
Then we add the planets into the disc at different initial
locations on circular orbits, and let them migrate. The plan-
ets do not accrete mass. In case of Kepler-47, we add three
planets, initially at 7.5, 14 and 20 abin. The two inner plan-
ets each have a mass of 25 M⊕ while the outer planet has
5 M⊕. The masses of the planets lie within the error bars of
the newly discovered 3 planet system in Kepler-47 (Orosz
et al. 2019). As the additional outer planet has much smaller
mass, and remains further out in the disc, it has little effect
on dynamical evolution of the two inner planets, and is not
discussed here any further. Even though in Kepler-413 only
one planet is observed, we embed two planets at 9 and 11.5
abin each with 0.21 MJup which is identical to the observed
planet mass in Kepler-413 (Kostov et al. 2014).
3. Forming a stable 1:1 resonance
In both cases, the planets migrate toward the stable parking
position near the inner cavity of the disc, where is first
reached by the inner planet, blue curves in Fig. 1. Such
low mass planets do not open a complete gap, hence their
orbits align with the cavity and gain in eccentricity (Kley
et al. 2019). The second planet, starting further out, takes
more time to reach the cavity and acquire eccentricity. This
Fig. 1: Time evolution of the orbital elements of the inserted
planets in the Kepler-413 and -47 system while embedded in
a disc, that was simulated for 20 000 Tbin prior to insertion.
perature, see Kley et al. (2019). They found the disc aspect
ratios, h, between 0.04 and 0.05, depending on the binary
system and disc parameters. Here, we choose h = 0.04, for
the viscosity parameter α = 0.001, and a disc mass of 1%
of the binary mass. The orbital parameter of the two mod-
elled systems, Kepler-413 and -47, are listed in the upper
block of Table 1.
The simulations presented here are performed using the
same approach and method as described in Thun & Kley
(2018) and Kley et al. (2019). We use cylindrical coordi-
nates centred on the barycentre of the binary and a radially
logarithmic grid with 684 × 584 cells, that stretches from
1 to 40 binary separations, abin. This domain size and res-
olution has shown to be sufficient for such type of studies
(Thun & Kley 2018). For the n-body integrator we use a 4th
order Runge-Kutta scheme, which is of sufficient accuracy,
given an average time-step size of 2 × 10−3 Tbin.
First, we simulate the discs without planets to their con-
vergent states which are reached after about 20 000 binary
orbits, Tbin, and then embed the planets. Simulations of cir-
cumbinary discs converge to a disc with a large, eccentric,
and precessing inner cavity. The mean orbital parameters
Article number, page 2 of 5
Anna B. T. Penzlin, Sareh Ataiee and Wilhelm Kley: 1:1 orbital resonance of circumbinary planets
In Kepler-47, whose planets are less massive than the for-
mer, the cavity almost maintains its previous precession
rate which is 2600 Tbin. The presence of these small planets
does not change significantly the cavity size and eccentricity
for either system.
The orbital dynamics of the two planets, engaged in the
1:1 resonance while still embedded in the disc, is displayed
in Fig. 3. It shows that the difference in mean longitude be-
tween the two planets never drops below 20◦. So they never
have a close encounter and do not collide, but are rather en-
gulfed in a periodic motion of a 1:1 resonance. In both cases
the planets move on horseshoe orbits around the Lagrange
points L4 and L5. In the case of Kepler-413 the motion is
quite regular with a libration period of about 1330 Tbin. In
the case of Kepler-47 the motion is more irregular and a
planet moves multiple time around one Lagrange point be-
fore switching over to the other. We suspect that the reason
for this irregular behaviour is the lower mass of the plan-
ets and stronger perturbed disc compared to Kepler-413.
In such a condition, the disc perturbation can give larger
random acceleration to the planets and send them back and
forth to the horseshoe and Lagrange points.
To check the long-term stability of these orbits in the
absence of the disc, we extracted the final positions and
velocities of the planets and stars, and used them as the
initial condition for a pure n-body simulation containing
the two stars with the planets (n = 4). In Fig. 4, we show
the evolution of semi-major axis and eccentricity over a
time span 22 000 Tbin. In both systems, the orbits are stable
and the planets remain in the 1:1 resonance. The orbits
retain their previous eccentricity (with disc) and change
periodically their positions upon close approach, where the
planet on the inner position switches to the outer and vice
versa, as is typical for horseshoe orbits.
In the pure n-body case, the orbits are much more reg-
ular than in the case with the disc, and in both systems the
planets are on regular, smooth horseshoe orbits, see Fig. 5.
Without the disc, the radial excursions of the planets are
similar to the case with the disc present. The libration pe-
riod is 1920 Tbin for Kepler-47 and 1055 Tbin for Kepler-
413. With the disc present, these periods were longer, for
example by 300 Tbin in the case of Kepler-413, because in-
teraction with the disc influences the orbital properties of
the embedded planets. During one libration period in the
n-body simulation, the planets approach each other only
up to a mean longitude difference between the planets of
about 25◦, shown in Fig. 5, ensuring stability.
Without the disc, the aligned planetary orbits precess
around the binary at a slower rate than with a disc present.
This is expected as the precession of the planets in the disc
was aligned with the inner cavity. In the n-body simulation
we find precession periods of 4245 Tbin for Kepler-47 and
2275 Tbin for Kepler-413. These periods correspond to that
of a single planet with the mean orbital elements of the
planet pair.
4. Discussion & Conclusion
Using a migration scenario for two planets in a circumbi-
nary disc we have shown that low eccentric binary systems
like Kepler-47 and -413 are able to host two small, equal-
mass planets in a stable coorbital resonance where the plan-
ets are on horseshoe type orbits with an average eccentric-
Article number, page 3 of 5
Fig. 3: Dynamical structure of the 1:1 resonances while the
planets are still embedded in the disc, on the left Kepler-413
and on the right Kepler-47. Top: Semi-major axis difference
of the planets against difference in mean longitude. The
colour changes linearly from blue to yellow over the time
span displayed in the lower panels. The red dots indicate
the location of the Lagrange points L4 and L5 in case of
circular motion. Bottom: Angular distance between the
planets vs. time.
process takes longer for Kepler-47, as lower mass planets
migrate slower, and the outer planet is initialised further
away.
As the parking position of planets around binaries is
determined by the disc's cavity edge, the incoming second
planet reaches the same orbit as the first. In our cases this
leads to capture in 1:1 orbital resonant configurations. As
shown in Fig. 1, this happens after ∼ 5000 Tbin for Kepler-
413 and ∼ 13 000 Tbin for Kepler-47. During the capture
process the two planets show initially larger excursions in
semi-major axis and eccentricity. During this process the
planets do not collide with each other, as the minimum
distance between them is ∼ 10−2 abin. The excursions are
damped by the action of the disc but the orbits remain ec-
centric. After the capture, we continue the hydrodynamical
simulations for more than 10 000 Tbin and the final orbital
configuration of the planets remains stable.
The shape of the orbits in relation to the disc's structure
in this phase of the simulations is displayed in Fig. 2, where
snapshots of the configurations are plotted at around the
capture time (left panels) and in equilibrium, at 20 000 Tbin
after planet insertion (right panels). The orbits lie near the
disc's inner edge, are more circular than the cavity but still
have an eccentricity of ∼ 0.25 − 0.30. They begin to align
just before capture, and their precession rate and that of
the cavity match exactly as soon as the planets arrive in
their stable parking positions. In the subsequent evolution,
the orbits remain fully aligned with a maximum libration
angle of 6◦. For Kepler-413, the precession rate of the cav-
ity is lowered to 1900 Tbin after introducing the planets.
A&A proofs: manuscript no. 1to1resonances
Fig. 4: Evolution of semi-major axes and eccentricities of the two planets orbiting Kepler-413 and -47 for the pure n-body
case without disc.
Having found possible resonant states for only two cases
of the Kepler sample, it is presently not known for which
orbital parameters of a binary star system stable coorbital
configurations can exits. Our investigated systems both
have very low ebin, 0.02 and 0.04, which appears to be an
important factor for the stability of such systems. Using
a comparable set-up for Kepler-35 (ebin = 0.14), we did
not find any coorbital resonant orbits but one of the plan-
ets was ejected after an interaction between the two plan-
ets. In the case of the highly eccentric binary Kepler-34
with ebin = 0.52 both planets were ejected after a close
encounter.
In addition to our equal planet mass studies we ran
a model with a different planet-planet mass ratio for the
Kepler-47 case. In this simulation, the mass of the inner
planet was about 20% higher than the outer one with oth-
erwise identical initial conditions. The system evolved again
into a stable coorbital configuration. However, in this case
the planets ended up in a tadpole type orbit around each
other with a mean eccentricity of about 0.28. Further ex-
ploration of the parameter space using different binary ec-
centricities and planet masses will be the topic of future
work.
The disc surrounding appears to be essential for the
trapping mechanism of the planets. We also tested the
Kepler-47 system with a less viscous disc α = 10−4. In
such a low viscosity disc, the planets are able to open a
gap, which leads to joint inward migration where the outer
planet pushes the inner planet inward, while exciting the
eccentricity of the inner planet. Thereby, the inner planet
gets ejected as soon as the outer planet migrated closer than
two binary separations to the inner planet. In the case of
massive planets, masses ≥ 1
3 MJup, both planets reach sta-
ble orbits at different semi-major axes. Therefore massive
Fig. 5: The difference in mean longitude of the two planets
for the pure n-body case without a disc. The colour changes
linearly with time over 1600 years or 57 900 Tbin for Kepler-
413 and 78 500 Tbin for Kepler-47.
ity of about 0.25 to 0.30. The orbits are aligned and precess
slowly around the binary.
These properties are different from coorbital resonances
observed so far. In the Saturnian system, Janus and
Epimetheus are on nearly circular orbits and in the case of
Trojans, very small objects are in resonance with a much
larger planet whose orbit is nearly circular. However, in the
latter case the small asteroid can be on an eccentric and in-
clined orbit which changes the resonant structure (Namouni
et al. 1999). In our case, the resonant planet pair orbits a
binary star on quite eccentric orbits.
Article number, page 4 of 5
Anna B. T. Penzlin, Sareh Ataiee and Wilhelm Kley: 1:1 orbital resonance of circumbinary planets
planets and inviscid environments may inhibit formation of
coorbital resonances.
Compared to the observations, our planets are located
farther away from the stars, which is the consequence of
their trapping at the parking position away from the cav-
ity edge. The more massive the planets are, the closer they
migrate towards the edge of the cavity. Also smaller disc
viscosities and aspect ratios bring the cavity edge closer to
the stars1. However, in such cases, the planets are more vul-
nerable to the gap-opening which, as explained, can prevent
them from getting into 1:1 resonance. Hence, we do not ex-
pect the coorbitals form very close to the stars such that the
stars' perturbation can affect them. To analyze the general
stability of such coorbitals closer to the stars and check if
they are similar to single planet orbits, will require exten-
sive future parameter studies for example through n-body
simulations.
After having shown that stable coorbital resonances
with moderate orbital eccentricities around binary stars are
a natural outcome of a convergent migration process, it will
be worthwhile to search for such systems in the Kepler data.
A recent study suggests that the single star TOI-178 is or-
bited by a coorbital planet pair, and an additional planet
(Leleu et al. 2019b).
Acknowledgements. We thank Zsolt Sandor for fruitful discussions.
Anna Penzlin was funded by grant KL 650/26-1 of the German Re-
search Foundation (DFG). The authors acknowledge support by the
High Performance and Cloud Computing Group at the Zentrum für
Datenverarbeitung of the University of Tübingen, the state of Baden-
Württemberg through bwHPC and the German Research Foundation
(DFG) through grant no INST 37/935-1 FUGG. All plots in this paper
were made with the Python library matplotlib (Hunter 2007).
References
Beaugé, C., Sándor, Z., Érdi, B., & Süli, Á. 2007, A&A, 463, 359
Cresswell, P. & Nelson, R. P. 2006, A&A, 450, 833
Cresswell, P. & Nelson, R. P. 2009, A&A, 493, 1141
Hunter, J. D. 2007, Computing In Science & Engineering, 9, 90
Kley, W., Thun, D., & Penzlin, A. B. T. 2019, A&A, 627, A91
Kostov, V. B., McCullough, P. R., Carter, J. A., et al. 2014, ApJ, 784,
Leleu, A., Coleman, G., & Ataiee, S. 2019a, arXiv e-prints,
14
arXiv:1901.07640
Leleu, A., Lillo-Box, J., Sestovic, M., et al. 2019b, A&A, 624, A46
Leleu, A., Robutel, P., & Correia, A. C. M. 2018, Celestial Mechanics
and Dynamical Astronomy, 130, 24
Michalodimitrakis, M. & Grigorelis, F. 1989, Journal of Astrophysics
and Astronomy, 10, 347
Mignone, A., Bodo, G., Massaglia, S., et al. 2007, ApJS, 170, 228
Namouni, F., Christou, A. A., & Murray, C. D. 1999, Phys. Rev. Lett.,
Orosz, J. A., Welsh, W. F., Haghighipour, N., et al. 2019, AJ, 157,
Pierens, A. & Raymond, S. N. 2014, MNRAS, 442, 2296
Rodríguez, A., Correa-Otto, J. A., & Michtchenko, T. A. 2019, MN-
Thun, D. & Kley, W. 2018, A&A, 616, A47
Thun, D., Kley, W., & Picogna, G. 2017, A&A, 604, A102
Yoder, C. F., Colombo, G., Synnott, S. P., & Yoder, K. A. 1983,
83, 2506
174
RAS, 487, 1973
Icarus, 53, 431
1 These results will be published in a future study.
Article number, page 5 of 5
|
Subsets and Splits